Vous êtes sur la page 1sur 950

A propos de ce livre

Ceci est une copie numérique d’un ouvrage conservé depuis des générations dans les rayonnages d’une bibliothèque avant d’être numérisé avec
précaution par Google dans le cadre d’un projet visant à permettre aux internautes de découvrir l’ensemble du patrimoine littéraire mondial en
ligne.
Ce livre étant relativement ancien, il n’est plus protégé par la loi sur les droits d’auteur et appartient à présent au domaine public. L’expression
“appartenir au domaine public” signifie que le livre en question n’a jamais été soumis aux droits d’auteur ou que ses droits légaux sont arrivés à
expiration. Les conditions requises pour qu’un livre tombe dans le domaine public peuvent varier d’un pays à l’autre. Les livres libres de droit sont
autant de liens avec le passé. Ils sont les témoins de la richesse de notre histoire, de notre patrimoine culturel et de la connaissance humaine et sont
trop souvent difficilement accessibles au public.
Les notes de bas de page et autres annotations en marge du texte présentes dans le volume original sont reprises dans ce fichier, comme un souvenir
du long chemin parcouru par l’ouvrage depuis la maison d’édition en passant par la bibliothèque pour finalement se retrouver entre vos mains.

Consignes d’utilisation

Google est fier de travailler en partenariat avec des bibliothèques à la numérisation des ouvrages appartenant au domaine public et de les rendre
ainsi accessibles à tous. Ces livres sont en effet la propriété de tous et de toutes et nous sommes tout simplement les gardiens de ce patrimoine.
Il s’agit toutefois d’un projet coûteux. Par conséquent et en vue de poursuivre la diffusion de ces ressources inépuisables, nous avons pris les
dispositions nécessaires afin de prévenir les éventuels abus auxquels pourraient se livrer des sites marchands tiers, notamment en instaurant des
contraintes techniques relatives aux requêtes automatisées.
Nous vous demandons également de:

+ Ne pas utiliser les fichiers à des fins commerciales Nous avons conçu le programme Google Recherche de Livres à l’usage des particuliers.
Nous vous demandons donc d’utiliser uniquement ces fichiers à des fins personnelles. Ils ne sauraient en effet être employés dans un
quelconque but commercial.
+ Ne pas procéder à des requêtes automatisées N’envoyez aucune requête automatisée quelle qu’elle soit au système Google. Si vous effectuez
des recherches concernant les logiciels de traduction, la reconnaissance optique de caractères ou tout autre domaine nécessitant de disposer
d’importantes quantités de texte, n’hésitez pas à nous contacter. Nous encourageons pour la réalisation de ce type de travaux l’utilisation des
ouvrages et documents appartenant au domaine public et serions heureux de vous être utile.
+ Ne pas supprimer l’attribution Le filigrane Google contenu dans chaque fichier est indispensable pour informer les internautes de notre projet
et leur permettre d’accéder à davantage de documents par l’intermédiaire du Programme Google Recherche de Livres. Ne le supprimez en
aucun cas.
+ Rester dans la légalité Quelle que soit l’utilisation que vous comptez faire des fichiers, n’oubliez pas qu’il est de votre responsabilité de
veiller à respecter la loi. Si un ouvrage appartient au domaine public américain, n’en déduisez pas pour autant qu’il en va de même dans
les autres pays. La durée légale des droits d’auteur d’un livre varie d’un pays à l’autre. Nous ne sommes donc pas en mesure de répertorier
les ouvrages dont l’utilisation est autorisée et ceux dont elle ne l’est pas. Ne croyez pas que le simple fait d’afficher un livre sur Google
Recherche de Livres signifie que celui-ci peut être utilisé de quelque façon que ce soit dans le monde entier. La condamnation à laquelle vous
vous exposeriez en cas de violation des droits d’auteur peut être sévère.

À propos du service Google Recherche de Livres

En favorisant la recherche et l’accès à un nombre croissant de livres disponibles dans de nombreuses langues, dont le frano̧ais, Google souhaite
contribuer à promouvoir la diversité culturelle grâce à Google Recherche de Livres. En effet, le Programme Google Recherche de Livres permet
aux internautes de découvrir le patrimoine littéraire mondial, tout en aidant les auteurs et les éditeurs à élargir leur public. Vous pouvez effectuer
des recherches en ligne dans le texte intégral de cet ouvrage à l’adresse http://books.google.com
This is a reproduction of a library book that was digitized
by Google as part of an ongoing effort to preserve the
information in books and make it universally accessible.

https://books.google.com
The University
of Michigan
transportation
Cibraiy

JNIVERSITY OF MICHIGAN
LIBRARIES

NOV 6

DEPOSITED BY THE
UNITED STATES OF AMERICA
Mechanics of Pneumatic Tires
Editor
Samuel K. Clark
University of Michigan
Ann Arbor, Michigan 48109

U.S. Department of Transportation


National Highway Traffic Safety Administration
Washington, D. C. 20590
Abstract
The pneumatic tire has been an integral part of automotive transportation almost since its
inception, yet it remains a product whose characteristics are not easily predictable or com
prehensible by conventional engineering techniques. This treatise is an attempt to provide a
rational descriptive and analytical basis for tire mechanics. Chapters of this book are con
tributed by active research workers in the fields of rubber and textile properties, friction, ma
terial properties, tire stress problems, tire design and construction, vehicle skid and handling,
and tire mechanical properties.

Key words: Friction; rubber, skid; tires; tire cord; tire contact; tire stress; tire structure;
vehicles.

For gale by the Superintendent of Documents, U.S. Government Printing Office


Washington, D.C. 20402
Preface
Shortly after the appearance of the first edition of this book it became
clear that a revision of it was needed, partly because the rapid develop
ment and acceptance of the radial tire in the United States dictated in
creased emphasis on that type of construction, and partly because certain
important topics, particularly wear, were not covered adequately. Further,
demand for the book quickly outran the available supply so that a new
printing seemed warranted.
Recognizing this need, the United States Department of Transportation
commissioned a thorough revision of the text in mid 1 975. Due to the
pressure of other commitments on some authors the final manuscript has
only now been assembled, a delay for which the Editor takes full responsi
bility.
I would like to acknowledge the efforts of the various authors of the
chapters of the revised edition, listed in the Contents, who have contrib
uted so effectively to the revision of this manuscript. In addition, I would
like to acknowledge the generous assistance of Dr. C. F. Barson, of the
Tire Technical Group, Dunlop Ltd., Birmingham, England, and to Dr. Jo
seph Walter of the Firestone Tire and Rubber Co., Akron, Ohio, both of
whom contributed substantially to the manuscript over and above any
credit they have received as authors of the various sections.

Samuel K. Clark
Editor

in
CONTENTS
"MECHANICS OF PNEUMATIC TIRES"

Chapter 1. RUBBER STRUCTURE AND PROPERTIES


S. D. Ciehman, formerly with
Goodyear Tire and Rubber Company, Akron, Ohio 44316
(Retired)

Chapter 2. TIRE CORD AND CORD TO RUBBER BONDING


T. Takeyama, J. Matsui and M. Hijiri
Tony Industries, Tokyo, Japan

Chapter 3. CORD REINFORCED RUBBER 123


J. D. Walter
Central Research Laboratories
Firestone Tire and Rubber Company, Akron, Ohio 44317
S. K . Clark
Department of Applied Mechanics and Engineering Science
University of Michigan, Ann Arbor, Michigan 48109

Chapter 4. STRUCTURE OF THE PNEUMATIC TIRE 203


V. E. Gough
formally
Dunlop Limited, Birmingham, England
(now deceased)
Chapter 5. CONTACT BETWEEN THE TIRE AND ROADWAY 249
Alan Browne
General Motors Research Laboratories, Warren, Michigan
K. C. Ludema
Department of Mechanical Engineering
University of Michigan, Ann Arbor, Michigan 48109
S. K. Clark
Department of Applied Mechanics and Engineering Science
University of Michigan, Ann Arbor, Michigan 48109

Chapter 6. TIRE TRACTION AND WEAR 365


A. Schallamach, formerly with
Malaysian Rubber Producers Research Association
Brickendonbury, Herts., England
(Retired)
K. Grose h
Uniroyal Inc., 5100 Aachen 1, Germany

Chapter 7. TIRE STRESS AND DEFORMATION _ 475


R. A. Ridha
Firestone Central Research Laboratory
Firestone Tire and Rubber Company, Akron, Ohio 44317
S. K. Clark
Department of Applied Mechanics and Engineering Science
University of Michigan, Ann Arbor, Michigan 48109
Chapter 8. MEASUREMENT OF TIRE PROPERTIES 541
H. van Eldik Thieme and A. J. Dijks
Vehicle Research Laboratory
Technical University, Delft, Netherlands
Stephen Bobo
U.S. Department of Transportation, Transportation Systems Center
Cambridge, Massachusetts 02142
Chapter 9. ANALYSIS OF TIRE PROPERTIES 721
H. Pacejka
Vehicle Research Laboratory
Technical University, Delft, Netherlands

Chapter 10. PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 871


J. R. Beatty
B. F. Goodrich Research Laboratories
Brecksville, Ohio 44141
Chapter 1
RUBBER STRUCTURE AND
PROPERTIES
S. D. Gehman1

1.1.1. Introduction 2
1.1.2. Composition of tire compounds 3
1.1.3. Technical evaluations of physical properties of tire com
pounds 5
Stress-strain properties and evaluation of cure 5
Aging tests 8
Tear tests 9
Hardness 10
Dynamic tests _ 10
Flex cracking 10
Rubber abrasion tests 10
1.1.4. Rubber elasticity 13
Thermodynamic aspects 13
Molecular picture: elasticity of a rubber molecule 15
Elasticity of the molecular network 17
Strain energy representation of rubber elasticity 19
1.1.5. Rubber viscoelasticity 20
Molecular and model concepts of rubber viscoelasticity 21
Viscoelastic relations between creep, stress relaxation, and
complex dynamic modulus 22
Superposition principle 23
Time-temperature superposition principle 24
Dynamic properties 25
Energy losses in tires 27
1 . 1 .6. Reinforcement of rubber with carbon black 29
References 33
1 Formerly with The Goodyear Tire & Rubber Company. Akron. Ohio 44316.
(Retired. Present address: 214 Kemlworth Drive, Akron. Ohio 44313.)
MECHANICS OF PNEUMATIC TIRES

1.1.1. Introduction
Pneumatic tires usually contain a variety of rubber compositions, each
designed to contribute some particular factor to overall performance.
Rubber compounds designed for a specific function will usually be similar
but not identical in composition and properties, although in some cases
there can be significant differences between compounds in tires of various
types. The guiding principle in development of rubber compositions for
tires is to achieve the best balance of properties for a particular type of tire
service. Since a tire is a mechanical structure, a rubber component should
be judged on how it functions in the system rather than on its individual
properties or performance capabilities. Thus a rubber compound which
did not adhere well to other tire components, or which required vastly dif
ferent vulcanization conditions than other parts of the tire, could be use
less in the tire even though it had excellent strength and other mechanical
properties. Tire performance is the result of skill and experience in pro
ducing a mechanically harmonious structure of rubber compounds, fabric
and adhesive, beads and other components which work together to give
optimum service.
The principal functions of the rubber compositions in a tire are fairly
obvious. The tread compound must provide wear resistance and be tough
and resilient to minimize cuts, tears, and cracks, as well as to protect the
tire body from bruising impacts. Low mechanical hysteresis loss in the
tread is desirable since lower tire operating temperatures are advanta
geous. Good friction properties of the tire tread for all driving conditions
are, of course, very important. In some cases optimum tread properties are
obtained by using a cushioning compound between tread and tire body as
additional protection against fabric bruises, thus making a "double layer"
tread. This cushion can also serve, especially in retreading, as a bonding or
transition layer between tread and body compounds. Intermediate hard
ness properties between those of tread and body are usually used in the
cushion or breaker under the tread.
Tire body or carcass rubber compounds must form strong bonds to the
adhesive-coated fabric. Their strength and durability should be adequate
to insulate the cords and hold them in their paths. The rubber must, how
ever, be soft enough to permit a slight change of cord angles when the tire
is flexed. The body rubber serves as insulation between the fabric plies.
Outstanding fatigue resistance is required of body compounds in order to
withstand cyclic deformation. It is essential that they retain adequate
physical properties and durability at the internal tire temperatures gener
ated in service. Hence, low mechanical energy losses are needed for body
compounds. There may be gradations in the properties of body com
pounds, with hardness usually diminishing somewhat from tread to cush
ion to top plies to inner plies.
In tubeless tires, a liner or coating on the inside ply retards diffusion of
inflating air into the fabric, and protects against ensuing ply separations.
Rubber compositions around the wire bead are called bead insulation,
and give it geometric stability, shape it to fit the rim, and provide firm an
chorage for the cords.
RUBBER STRUCTURES AND PROPERTIES 3

Finally, sidewall compounds must be especially durable in flexing and


weathering, and be scuff and impact resistant to protect the body from
curb impacts. The sidewall may include a decorative compound as a sur
face layer.
Processing requirements impose additional restraints on rubber compo
sitions suitable for tires, so that the rubber technologist must consider
many factors in compounding a rubber for a specific use.

1.1.2. Composition of Tire Compounds


Ingredients [1]2 in tire compounds can be classified as: (1) the rubber
which may be a single polymer or a blend of polymers [2, 3] and, with
high molecular weight polymers, may include an extending oil [4-6]; (2)
fillers, principally various types of highly-developed carbon blacks [7-9];
(3) relatively small additions of softeners, plasticizers, or reclaim rubber
which serve principally as processing aids; (4) the chemical vulcanization
system [10-12], which is likely to include two accelerators [13], sulfur, and
a small amount of zinc oxide; (5) chemical protective agents, known as
antioxidants and antiozonants [14, 15].
With such a wide variety of ingredients, the important mechanical prop
erties for a given tire compound can usually be obtained from a number of
different compositions. For example, modulus and hardness can be con
trolled by varying either the amount of carbon black, the amount of ex
tender-oil or softener, the fineness and structure of the black, or the num
ber of molecular crosslinks introduced during vulcanization. Thus even
small advantages in cost, performance, and processing, which may only
become apparent with extensive testing or service experience, become im
portant factors in compound selection. No tire compound is ever final but
is always subject to changes as test results and experience accumulate.
In the United States, in 1968, about 75 percent of the rubber being used
in tires and tire products was synthetic [16], and of this about 73 percent
was SBR (styrene-butadiene copolymers) [17], 21 percent was stereo-
elastomer [17] and about 5 percent was butyl [18] or chlorobutyl rubber
[19]. The latter is used for liners and inner tubes because of low air per
meability and resistance to oxidative deterioration. The stereo rubber, cis
1,4-poly-butadiene, has come into wide use as a compounding ingredient
which can improve durability of tire body compounds, can reduce groove
cracking and can increase abrasion resistance of tire treads.
Airplane tires have been the last stronghold of completely natural rub
ber tires, but synthetic rubber is now beginning to appear in them also.
Of the 30 or so standard types of carbon black, the HAF (High Abra
sion Furnace) and ISAF (Intermediate Super Abrasion Furnace) types
with their HS (High Structure) variants are currently used most exten
sively in tire treads. There is now a trend toward high structure blacks in
treads, i.e., from ISAF black to HAF-HS. Relative use of HAF and ISAF
in 1967 was about 68 percent and 32 percent respectively [20]. The high
structure blacks are reported to give somewhat better wear and resistance
to cracking and cut growth, along with improving processing of polymers
used in treads. A physical description of these blacks is given in table
1.1.3 of this chapter.

2 Figaro in brackets indicate the literature references at the end of this chapter
MECHANICS OF PNEUMATIC TIRES

TABLE 1.1.1. Tread-typeformulations


Example ' Example
A [23] B[24]
OE-SBR* 89.38 96.00
BR* 35.00 30.00
Carbon black "70.00 +JO.OO
Zinc oxide 3.00 4.00
Stcaric acid 2.00 2.00
'1.00 1.50
19.63 24.00
M.05 *1.20
^02
Sulfur 1.86 2.00
222.94 230.70
Cure 40 min/
293°F
'Oil-extended sytrene-butadiene copolymer, 100 parts polymer,
37.5 parts oil
* cu-polybutadiene
c HAF or ISAF types
" HAF types
•N-Cyclohexyl-N-phenyl-/>-phenylenediamine
-''Aromatic secondary a mine type
*N-Oxidiethylene-2-benzothiazylsulfenamide
* Sulfenamide type
' Diphenyl guanidine
Larger particle size carbon blacks, such as SRF, GPF, and FEF types
are usually preferred for carcass compounds.
Representative tread and carcass compound recipes can be found in the
rubber technical literature [2, 3] as can cushion or undertread formula
tions [21], wire bead insulation [22], and liner recipes [19].
For illustration a few recipes are given which are representative of tire
compounds.
Simpler standard formulations are available for test and material speci
fication purposes [25], and these, or similar ones, are often preferred for
research studies on physical properties of rubber compounds.

TABLE 1.1.2. Carcass-typeformulations [2]


Truck tire Passenger
body tire body
compound compound
Smoked sheet 25.00 25.00
Pale crepe 25.00 25.00
cii-BR 50.0 50.00
FEF Carbon black 40.00 35.00
Medium processing oil.. 8.00
8.00
Rosin oil 2.00
2.00
Wing-Stay 100.. 1.00 1.00
Stearic acid 2.00 2.00
A max #1 .80 .80
Zinc oxide 5.00 5.00
Sulfur 2.50 2.50
Optimum cure.. 40 min/ 8 min/
275°F
RUBBER STRUCTURES AND PROPERTIES 5

1.1.3. Technical Evaluations Of Physical Properties Of Tire


Compounds .
Evaluation of the physical properties of rubber compounds for technical
purposes proceeds by subjecting them to a battery of standard laboratory
tests [26, 27]. Occasionally individual laboratories may deviate from these
standard procedures, and may also use special tests and testing devices
which they have found to be advantageous. However, the standard tests
for rubber in the United States represent many years of testing experience
[4]. International standardization of rubber tests is also well advanced and
under constant development [28]. It should be realized that these standard
tests, although they fulfill practical purposes in rubber technology, often
involve combinations of several basic properties of rubber and hence are
of limited general use. They usually do not furnish a very fundamental de
scription of the physical properties, such as might be desired for research
purposes.
Here we discuss very briefly the most important of these test procedures,
and examine a few typical results when they are applied to tire com
pounds.
Stress-strain properties and evaluation of cure. The physical properties of
any rubber compound depend upon the state of cure [10], that is, upon
how far the chemical vulcanization reactions have been carried. Vulcan
ization introduces chemical crosslinks or bonds between the long chain
polymer molecules. This crosslinked network is decisive for the physical
properties and is determined, for a given rubber compound, by vulcaniza
tion time and temperature. The traditional rule of thumb is that the vul
canizing time to reach a given level of a property, such as static modulus,
is halved if vulcanization temperature is raised 18 F, and vice versa. Al
though this rule is still often adequate, a more precise description of the
time-temperature dependence of vulcanization requires determination of
the activation energy of vulcanization for each rubber compound [10, 29,
30].
The effect of cure on the physical properties of a rubber compound is
usually determined by vulcanizing a series of test sheets for different times
at the same temperature [25]. This may, of course, also be done at several
temperatures if more thorough tests are desired. Dumbbell-shaped test
specimens are cut from the sheets and static stress-strain curves taken [31].
Rubber stress-strain curves under static conditions are concave toward
the load axis, i.e., strain hardening, except for a short portion near the ori
gin. The concavity is accentuated by the occurrence of stress-induced crys
tallization in natural rubber at higher elongations [32, 33]. There is no
yield point before failure, as is usual with metals. The curves hi figure
1.1.1. represent approximately the effect of cure time on static stress-strain
properties for a range of cures of a tread-type compound. Stress and ten
sile strength are calculated using the original cross-section area. In rubber
technology, this stress is called modulus and is designated for a specific
elongation, so that the 300 percent modulus for a rubber compound is the
stress required to extend a strip to four times its original length. Volume
changes are negligible [34], when rubber is strained, so that Poisson's ratio
is assumed to be one-half.
Static modulus provides a convenient parameter to assess the temper
ature range in which elastomeric properties are exhibited. While this var
ies somewhat with individual polymers, in general the modulus of an
MECHANICS OF PNEUMATIC TIRES

CURE
CURVE'MINUTES AT 3IO*F

200 400 600


STRAIN , %

FIGURE 1.1.1 Variation of stress-strain curves with cure; data from reference [108}.

elastomer varies with temperature as shown in figure 1.1.2. At low temper


atures a hard or glassy character is evident. As temperature is raised the
rubber passes through a transition region in which properties change rap
idly. Rubber properties prevail over a range of temperatures above the
transition temperature, and finally at yet higher temperatures viscoelastic
or flow properties become important and predominate. It should be noted
that properties other than modulus generally also show significant changes
in passing from the glassy to the rubbery state.
At high temperatures and for long times, the flow properties of rubber
are marked. Stress relaxation in this region can often be attributed to
oxidative degradation. In the elastomeric region there is relatively little
stress relaxation, and in this state the molecular network comes into equi
librium with applied stress, so that here the concept of modulus is valu
able.
As shown in figure 1.1.1., modulus or stress at a given elongation in
creases, and breaking elongation decreases, as cure advances. This is al
most invariably the case. Tensile strength usually goes through a maxi
mum, although this is not always observed in a range of test cures. In
some compounds a phenomenon known as reversion occurs and modulus,
tensile strength, and breaking elongation all decrease with over-curing.
Natural rubber is especially prone to this reversion. Stress-strain curves
such as those shown in figure 1.1.1. are widely used in tire compound de-
RUBBER STRUCTURES AND PROPERTIES

GLASSY JTBANSITON ELASTOMERIC VISCOELASTIC


"REGION" I^REGION REGION OR
FLOW REGION

TEMPERATURE

FIGURE 1.1.2 General variation of rubber modulus with temperature.

velopment to adjust cure rates and to insure that modulus, tensile strength,
and breaking elongation of a compound fall inside a desired range. For
tire tread compounds the usual tensile strength will be in the range 2500 to
4000 psi, 300 percent modulus in the range of 1000 to 1700 psi, and break
ing elongation in the range of 400 to 600 percent.
Although such discrete test cures are traditional in the rubber industry,
and are useful for evaluation of physical properties in relation to cure, var
ious instruments are also used which furnish a continuous record of mod
ulus or stiffness as cure progresses [30, 35]. Such data are very useful in
showing how modulus develops, whether or not it reaches a flat plateau
and whether or not there is reversion or decrease in modulus with over-
cure. Figure 1.1.3 illustrates these possibilities diagrammatically. The
character of such a curve is determined by the polymer and the vulcan
izing system [10]. The start of the curve is also significant, because it gives
a measure of processing safety, that is, an indication of time-temperature
conditions which the compound can endure in mixing, extrusion, etc.,
without excessive prevulcanization or "scorching".
Optimum or "best" cure for a rubber compound cannot in general be
uniquely defined, since it depends upon the type of service and the partic
ular property, or properties, which should be optimized [36]. The cure des
ignated "best" in figure 1.1.1 gives a balance of properties best suited by
experience for tire tread service. There is usually a range of cures for
which a vulcanizate would be expected to give about the same service per
formance, but large deviations from the best cure can be disastrous. Opti
mum cure time is usually found to be within the time required to obtain
90 to 95 percent of the rise in modulus, as shown by a curve such as in fig
ure 1.1.3.
When a tire is cured, the time-temperature history will vary throughout
the tire cross section since heat must flow into it from the mold and
through the internal bladder used to form the tire. The thermal diflusivity
of rubber is relatively low [37], so that the equivalent cure, i.e., minutes of
cure expressed through calculation at a given reference temperature, may
vary considerably between the outer surface and the interior point of least
cure, even for a passenger tire. It is desirable to formulate tire compounds
MECHANICS OF PNEUMATIC TIRES

MARCHING MODULUS
FLAT
REVERSION

CO

20 40 60 80
CURE TIME , MIN.
FIGURE 1.1.3 Diagram of continuous cure curves.
Curve A u representative for an SBR tread compound cured at 2SO°F; the term "marching modulus" indicates a slowly in
creasing modulus for long cure times.

to compensate for this as much as possible, and to use vulcanization sys


tems which provide "flat" cures, that is, cures such that physical properties
tend to reach a plateau as vulcanization progresses. It is important to have
a good, adequate cure at every point in the tire and yet to avoid excessive
cure at any point.
Tire cures are originally determined by incorporating thermocouples at
strategic points in the tire. Temperature is measured during vulcanization
as a function of time. The production curing cycle is gotten by analysis of
these time-temperature curves in relation to the vulcanizing characteristics
of the compounds. Cure times for passenger car tires may be on the order
of 20 minutes [38].
The state of cure of the rubber in a tire can probably be most conve
niently examined or studied by equilibrium swelling measurements in
benzene using small pieces of rubber cut from the tire [10]. The molecular
crosslinks formed during vulcanization limit the swelling.
Aging tests. Physical properties of tire compounds, especially stress-
strain properties, are also routinely examined in rubber laboratories after
the rubber has been exposed to one or more accelerated aging tests [39].
These are usually run at elevated temperatures in order to simulate deteri
oration in service over a long period of time. Several of the tests are oven-
aging tests under carefully controlled conditions, while others use a bomb
filled with air or oxygen under pressure to further accelerate degradation.
Figure 1.1.4 gives an example of the effect of oven aging on the stress-
strain curve of a tread compound. Although such results do not correlate
perfectly with aging deterioration in service, they can be very helpful, es
pecially with a background of experience, in anticipating whether or not a
compound will be satisfactory in this respect. These aging effects are quite
complicated [40, 41] as they depend on oxidative chemical reactions with
the polymer [42, 43]. Hence they are very dependent both on the
RUBBER STRUCTURES AND PROPERTIES 9

nature of the polymer and on antioxidants in the compound recipes. In


general, SBR is less sensitive to aging than natural rubber. The basic
mechanism of degradation in these two rubbers appears to be quite differ
ent, SBR tending to harden on aging and natural rubber to soften, reflect
ing, respectively, predominance of additional molecular crosslinks and
chain scission, i.e., cutting of chain molecules into smaller molecules.
Effectiveness of antiozonants in rubber compounds is usually evaluated
by outdoor exposure tests [44] and accelerated ozone tests [45], to promote
the characteristic cracks which result from action of ozone on stretched
rubber [46, 47].
Tear tests. Rubber tear tests [48] are designed to cause a high stress gra
dient at the end of a cut or notch in an angle or crescent shaped testpiece,
which is pulled in a testing machine. Figure 1.1.5 illustrates various types
of tear test specimens. Although tearing phenomena with rubber are most
important and revealing in regard to mechanisms of rubber failure [49,
50], technical tear tests have very limited practical significance, probably
because the notch effect is so complicated and difficult to control for rub
ber. Tear test values are reported as load per unit of specimen thickness.

OE SBR TREAD COMPOUND


BEST CURE
3000

2 2000

w
w
<r

1000

200 400 600


STRAIN, %

FIGURE 1.1.4 Effect of air-oven aging on stress-strain curve of the rubber compound of
Figure 1.1.1.
10 MECHANICS OF PNEUMATIC TIRES

An SBR tread compound with tensile strength about 3000 psi might rea
sonably show a crescent tear strength of about 250 lb/in. In general, tear
values tend to increase with greater tensile strength and breaking elonga
tion but to decrease with higher modulus.
Hardness. Rubber hardness is an important quality control parameter.
It is conveniently measured with a Shore A Durometer, 3 a pocket in
strument which has been standardized [51] but use of which often leaves
much to be desired in the way of precision. This may be improved by
mounting it in a rigid stand. The durometer uses a small, spring loaded in-
denter with a truncated conical point protruding from a flat base. When
indenter and base are pressed against the rubber, the resulting spring de
flection, which depends upon rubber hardness, is indicated by a pointer
with a scale graduated from 0 (no hardness) to 100 (no indentation). Shore
A hardness for rubber tread compounds is typically in the range of 50 to
65 units and for unfilled vulcanizates about 25 to 30.

3 Certain commercial products and instruments are identified in this book in order to specify adequately technicalJproce-
dures. In no case does such identification imply recommendation or endorsement by the National Bureau of Standarrds, nor
does it imply that the products or equipment identified are necessarily the best available for the purpose.

STRIP TEST SPECIMEN TROUSERS TEST SPECIMEN

CRESCENT TEST SPECIMEN PURE SHEAR TEST SPECIMEN

ANGLE TEST SPECIMEN RING TEST SPECIMEN

FIGURE 1.1.5. Tear test specimens.


RUBBER STRUCTURES AND PROPERTIES 11

There are a variety of other hardness test instruments for rubber and a
well-developed International Rubber Hardness Degree (IRHD) scale
which agrees approximately with the Shore A Durometer scale [52, 53].
Indentation of thick rubber obeys the classical elasticity analysis of Hertz
very well [54], and this gives a mechanism to relate elastic modulus to
hardness measurements.
Dynamic tests. Many different test procedures are available to measure
rubber stiifness and energy loss for relatively small cyclic deformations,
often over ranges of temperature and frequency. These evaluations are es
pecially pertinent for tire compounds because heat generation and temper
ature rise from rubber hysteresis losses are important factors in tire du
rability [55].
One of the oldest and still most widely used types of test for this purpose
is a pendulum rebound test, in which a pendulum is released from a fixed
height to strike a rubber block and then rebound [56]. Superiority of natu
ral rubber or synthetic cis-1,4-polyisoprene in this test is pronounced. Per
cent rebound for SBR tread compounds will usually be in the range of 52
to 62 percent while that for comparable natural rubber compounds may
be more than 70 percent. A falling ball instead of a pendulum is often
used in a rebound test.
Free vibration tests [57] and forced nonresonant and resonant vibration
tests are also found in great variety [58, 59]. These are used to measure dy
namic modulus,4 internal friction,4 and resilience4 of rubber compounds.
Reference [60] includes a useful table of storage4 and loss moduli4 for a
wide variety of rubbers and other materials.
Energy losses in rubber may be evaluated by measuring the temperature
rise when a block specimen is cyclically deformed in shear or compression
or both. These rubber testers are called flexometers [61]. There are many
limitations, however, on the usefulness and interpretation of such temper
ature rise data. Compression set of the rubber specimen after a flexometer
test is usually also reported and is used in compound evaluations, espe
cially for state of cure. Occasionally the tests are run to destruction.
Flex cracking. Initiation and growth of small cuts or tears in tire treads,
especially in the pattern grooves and in tire sidewalls, are so significant for
tire performance that much effort has been expended in developing labo
ratory tests to simulate and clarify these phenomena. Several testing ma
chines and procedures have been standardized for such tests [62]. The De-
Mattia machine, where a specimen with a transverse groove is cyclically
bent as shown in figure 1 . 1 .6, is probably the most familiar of these. Re
sults of such tests are sensitive to compounding factors such as type of
elastomer, state of cure, and protective agents in the recipe [63].
Rubber abrasion tests. Laboratory abrasion tests are used in tread com
pound development because they provide inexpensive, rapid screening of
experimental polymers and rubbers, in spite of the fact that correlations
with roadwear have limited success. Two methods are given in ASTM
Standards [64], the Dupont abrader where two rubber specimens are abra
ded against a revolving disk, and the Pico abrader in which the surface of
the specimen is rubbed by revolving tungsten carbide knives under care
fully controlled conditions. Much abrasion testing is also done with the
Goodyear angle abrader [65], in which a specimen in the shape of a small
rubber wheel is mounted at a slip angle and driven by a revolving abrasive
wheel. A variation of this is the Lambourn [66] abrader, in which slip of
4 Thoe lerou »re defined in the lection on dynamic properties.
12 MECHANICS OF PNEUMATIC TIRES

k
1"
140- 55mm

1 §
/"(RO.O94 «O.OOI in) 0 0
/ a*fi
2t s s
(0)

FIGURE 1.1.6
(•) DeMallia lestpicce (b) Mounting of DeMattia lestpiecc.

the rubber wheel specimen is controlled by an electromagnetic brake so


that tests can be run either at constant slip or constant transmitted power.
Developments with Lambourn-type abraders have become quite sophisti
cated in efforts to secure correlations with road tests on different road sur
faces and under different driving conditions [67].
It is widely recognized that the severity of abrasion, or rate of wear, is
an important factor in abrasion testing. Rankings of a series of compounds
may change if rate of wear is changed. This introduces obvious complica
tions in trying to correlate laboratory data with road wear, since severity
of wear on the road depends on many factors including road surface, fre
quency of curves, amount of rain, braking and many others. The severity
level of a laboratory abrasion test, that is, the amount of rubber removed
per unit distance traveled, should presumably be comparable to that in
road tests for best correlations. This may be very low, by American stan
dards, in the order of 100 miles/mil5 of wear normally and 50 miles/mil or
less for severe service. To reduce testing time, laboratory tests are usually
accelerated and have much larger wear rates than road tests.
'One mil equals .001 inch equals .0254 nun.
RUBBER STRUCTURES AND PROPERTIES 13

1.1.4 Rubber Elasticity


Thermodynamic aspects. Distinctive features of rubber elasticity are easy
deformability or low modulus, enormous deformations, and rapid recov
ery when deforming forces are released. There is also more sensitivity to
temperature than for many elastic materials. Figure 1.1.7 shows the dra
matic effect of low temperatures on relative modulus of unfilled vulcan-
izates of several polymers. For any elastomer there is a range of temper
atures over which transition occurs from a rubbery to a glassy state, as
shown in figure 1.1.2. This transition temperature range for SBR is from
-60°C to -40°C, which is about as high as can be tolerated for a general
purpose tire rubber. The curve for m-polybutadiene in figure 1.1.7 is com
plicated by crystallization, which starts to affect the warming curve at
about -95°C.
In the transition range from the glassy to the rubber state modulus falls
rapidly with increasing temperature, but further temperature increase re
sults in a slowly rising modulus. This rise was first observed in experi
ments by Gough, published in 1805, as a contraction when a rubber speci
men stretched by a weight was heated. Joule, about 50 years later, studied
thermoelastic phenomena exhibited by rubber and interpreted them in
terms of the new science of thermodynamics then being developed by Kel
vin. Treloar [72] gives a very good review of the thermodynamic funda
mentals of rubber elasticity. For a reversible process, the first and second
laws of thermodynamics provide,
dE=TdS (1.1.1)

100

111

-100 -80 -60 -40 -20


TEMPERATURE, «C
FIGURE 1.1.7 Effect of low temperatures on rubber modulus; data by ASTM method
1053-65, reference [4J.
14 MECHANICS OF PNEUMATIC.TIRES

in which E is internal energy of a system, T is absolute temperature, S is


entropy, and W is work done on the system. At once there is a difficulty
here because ordinary rubber deformations are not completely reversible.
It is necessary to take special measures with any test specimens, such as
solvent vapor treatments or prestretching them at an elevated temperature,
in order to secure reversible deformations.
If the tensile force on a rubber strip is ⨍, then the work done during an
isothermal displacement dl is, neglecting small volume changes,
dW=idl (1.1.2)
and, with eq (1.1.1)
SW] 8E_
- T (1.1.3)
SI 81 51

Equation (1.1.3) resolves the force into two terms. The first arises from
changes in internal energy and the second from entropy changes with
changes in length.
By differentiation of eq. (1.1.3), it follows that

(1.1.4)
ST SI

so that eq ( 1 . 1 .3) can be written


+ T\-L- (1.1.5)
«/ r \ST

Equation (1.1.5), in conjunction with eq (1.1.4), has been very important


for understanding rubber elasticity because it allows experimental evalua
tion of internal energy and entropy changes upon deformation. If the
equilibrium force exerted by a stretched rubber strip held at constant
length is measured and plotted as a function of temperature, the slope at
*/
any value of T is ST , which eq (1.1.4) shows to be the entropy change
per unit change in length for isothermal extension at T. The corresponding
SE\
internal energy change -35-] is given by the intercept of the tangent with
the zero Taxis. °' r
Careful experimental work of this type by Meyer and Ferri [68], An
thony, Caston, and Guth [69], Wood and Roth [70], Gee [71] and many
others showed that over a considerable temperature range stress was
closely proportional to absolute temperature, and this led to the conclu
sion that rubber elasticity resides principally in the entropy term of eq
(1.1.5). There is an entropy decrease on extension and an increase on re
traction, except:
a. at very low elongations, below about 10 percent where a so-called
thermoelastic inversion is observed due to thermal expansion obscuring
the entropy effect
b. at large elongations, where high orientation and crystallization may
occur.
RUBBER STRUCTURES AND PROPERTIES 15

Volume changes and internal energy effects, however, appear never to be


entirely absent. This entropy basis indicates that rubber elasticity must
have an entirely different molecular origin or mechanism than ordinary
elasticity, where stresses increase the internal energy by increasing molec
ular or atomic spacings.
Molecular picture: elasticity of a rubber molecule. The unique thermoe-
lastic behavior of rubber is related to molecular structure by the kinetic or
statistical theory of rubber elasticity. The theory provides a very satisfac
tory explanation of what might be called the mainspring of rubber elastic
ity, but it involves idealizations which have restricted its quantitative ap
plication to very carefully controlled equilibrium experiments with
suitable rubber compounds and limited elongations and temperatures. It
can only be regarded as semiquantitative for rubber in real applications.
Reasons for these deviations from the theory, however, are quite com
prehensible in light of what is known from many sources concerning the
molecular geometry and forces.
There is extensive evidence that a rubber is composed of long chain
molecules as shown in figure 1.1.8. The monomer repeating unit in the
chain molecule of r/.v-polyisoprene or natural rubber is
CH,
-CH,-C CH-CH2-
while in cw-polybutadiene it is -CH2-CH=CH-CH2-. In SBR, styrene
units amounting to about 23 percent by weight occur at random in the
polybutadiene chain. Such molecules are flexible by virtue of rotation
around the single bonds, except at low temperatures where packing be
comes too close or crystallization may occur for the first two rubbers. They
tend to assume haphazard or chance configurations because of thermal
agitation of their segments. Most of these configurations will be very
crumpled, so that a chain molecule can be extended by an external tensile
force provided that interaction with its neighbors is not too strong. As tem
perature is lowered, this interaction increases until the typical low modu
lus and rubber elasticity are no longer present. This temperature influence
is shown in figure 1.1.7.

ENTANGLEMENT^ J
</^r^ltfc- CROSSLINK

FIGURE 1.1.8 Diagram to illustrate concept of molecular structure of rubber.


16 MECHANICS OF PNEUMATIC TIRES

Chemical composition of the molecules may vary widely provided that


chain length, flexibility, and interactions all lie within ranges appropriate
for rubber elasticity. In order to secure mechanical stability in such a mo
lecular structure it is necessary to connect the chain molecules into a net
work by introducing chemical bonds or crosslinks between them during
vulcanization. The chemical nature of the crosslinks is not relevant for the
theory because they are idealized simply as network connections. The na
ture of the crosslinks is, however, technically important. The crosslinks
cannot be too numerous, or flexibility of the network will suffer, but their
number must be adequate to suppress plastic flow. To be more definite,
rubber chain molecules having 1000 to 2000 monomer units per chain
may very well have about 10 to 40 of these units crosslinked at random
along a chain in a vulcanizate.
History of the kinetic theory of rubber elasticity, which dates back to
about 1932, has been recounted by Treloar [72] and by Flory [73]. The
theory has been thoroughly examined both experimentally and analyti
cally and has been refined in many respects. Treloar gives an excellent ex
position of developments up to about 1958, and a review article by Krig-
baum and Roe [74] brings the subject up to recent years. Only a brief
recapitulation of a few of the main features of the theory will be given
here.
A single, long chain rubber molecule assumes random statistical config
urations to the extent permitted by hindered bond rotations, fixed valence-
bond angles,6 excluded volumes (since no two atoms can occupy the same
space simultaneously), and intermolecular forces. Some of these effects
can be accommodated in the theory, but in the first mathematical model
the molecule is assumed to undergo random thermal fluctuations among
all possible configurations of its n links, each of length l, just as if they
were freely orienting. Thus the problem of describing the configurations is
the random walk problem, and the configurations have analogies to the
Brownian motion of a particle.
The distance between the ends of a chain, r, is called the displacement
length, end-to-end distance, or simply the chain length. The distance mea
sured along the chain is the chain contour length.
The distribution of chain lengths is Gaussian provided they are not ex
tended more than about one-third of their fully extend length, nl.
This distribution is given by
P(r)dr = (4£3/V/2)r2 exp (- b2r2)dr. (1.1.6)
In eq (1.1.6) P(r) is the probability function for r and
b2 - (3/2)/(/i/2). (1.1.7)
The root mean square value of r is
(f1) l/2 = fo"». (1.1.8)
This shows at once that n must be large to account for rubber elongations,
since the ratio of the average unst retched chain end separation to the fully
extended chain length is n-1/2. Higher molecular weights thus favor greater
extensibility.
Equation (1.1.6) is for a freely orienting chain. For a real molecule,

6 For i paraffin chain molecule UK ingle between two adjacent bonds in the -C— C tincture i» 109.5°
RUBBER STRUCTURES AND PROPERTIES 17

where valence angles are retained and rotation may be hindered, it is nec
essary to introduce the idea of an equivalent random chain and an equiva
lent random link. Length of the equivalent random link depends upon the
chemical structure. Usually the equivalent link contains up to ten main
chain bonds. For polyisoprene, Treloar estimated that there were 0.76
monomer units per random link. If a rubber network chain molecule, i.e.,
a chain between crosslinks, contains about 80 monomer units, or 105
freely orienting links, the ratio of fully stretched length to average unst-
retched end-to-end distance will be (105)1/2 ≈ 10.2. This is adequate to ac
count for rubber elongations.
According to the familiar Boltzmann relation the entropy of a system is
proportional to the logarithm of the number of possible configurations.
Hence from eq ( 1 . 1 .6) the entropy 5 of a single chain molecule is
S-c-k-b2-r* (1.1.9)
in which c is an arbitrary constant and k is Boltzmann's constant. It is ap
parent from eq ( 1 . 1 .9) that the entropy decreases as r becomes larger, that
is, as the molecule is stretched. The work required to increase r to r + dr is

fdr=-T (1.1.10)

in which / is the stretching force on the molecule. From eq ( 1 . 1 .9) dS/dr =


- 2kb2r so that
f = 2kTVr. (1.1.11)

This is the average fluctuating statistical force exerted at the end of a


stretched molecule. It is proportional to absolute temperature and to the
value of r, the end-to-end distance, and it acts along the line of r.
Elasticity of the molecular network. A chain molecule reaching from one
crosslink to another is called a network chain. The Gaussian distribution
of eq (1.1.6) is assumed to apply to each network chain so that the entropy
change for deformation of the network can be calculated by summing the
entropy changes for all the chains in the network. In doing this the as
sumption is made that the deformation in affine, that is, the vector com
ponents of length of each chain are changed in the same ratio as the corre
sponding dimensions of the rubber specimen. Treloar [72] gives the
entropy change ∆S of the network due to deformation as
t + \l + \l - 3) (1.1.12)
in which N is the number of network chains per unit volume and λ1, λ2,
and λ3 are the principal extension ratios along the three mutually per
pendicular axes of strain for a pure homogeneous strain. The extension ra
tio is denned as the ratio of the deformed to the undeformed length.
If the deformation is not accompanied by any change in internal energy
the work of deformation, W, is — 7AS, so that
\l + \l - 3) - JG(A? + \l + \l - 3) (1.1.13)
W is the elastically stored free energy per unit volume and is known as the
stored energy function.
In eq (1.1.13) G is the modulus of rigidity which in this simpler version
18 MECHANICS OF PNEUMATIC TIRES

of the theory depends uniquely on chain molecular weight through the re


lation
G = NkT=pRT/Mr. (1.1.14)
Here ρ is the density of the rubber, R is the gas constant and Mc is the
number-average network chain molecular weight. The theory thus pro
vides a single elastic constant which is proportional to the degree of cross-
Unking.
To calculate the principal stresses from eq (1.1.13) it is assumed that
volume changes can be neglected so that
X,-X2-X,-1 (1.1.15)
Work done by the applied forces is
fld\l+f2d\2 + f,d\3 (1.1.16)
in which /„ /,, and /., are forces per unit initial unstrained area and act
along the principal axes. Comparing dW obtained by differentiating eq
(1.1.13) with </ W in eq ( 1 . 1 . 1 6), after eliminating X, from both by use of eq
(1.1.15), and equating the coefficients of dλ1 and dλ2 gives the general
stress-strain relations:
A!), (1.1.17)
Xi). (1.1.18)
Equations (1.1.17) and (1.1.18) can be written in terms of principal stresses
/„ fj, t3, denned as forces per unit area after straining, by use of relations
such as /, = f,X2X3, in the form
/, - f, - G(X? - XJ), (1.1.19)
t, - f3 = G{\\ - X|). (1.1.20)
These equations give only the difference between two principal stresses,
since eq (1.1.15) has introduced the indeterminancy of an arbitrary hydro
static stress. This may be recognized by writing the principal stresses [72]
as
f, = G\] +p\ ti = G\\+p; /3 = G\l +p. (1.1.21)
However, if one or more of the principal stresses is given, a unique solu
tion can be obtained for the other two stresses provided that the extension
ratios λ1, λ2, X, are known. For uniaxial extension /2 and /3 both vanish,
and λ2 = λ3 = X7W so that from eqs (1.1.17) and (1.1.18), the rubber stress-
strain relation is
/ = G(X-X-2). (1.1.22)
Treloar carried out experiments to verify eqs (1.1.19) and (1.1.20) using
simple extension, uniaxial compression, uniform two dimensional exten
sion and shear deformation on natural rubber gum vulcanizates. He found
that these equations using the single physical constant G provided a fairly
satisfactory first approximation to experimental results. However, devia
tions were observed both at moderate strains, where measured stresses
tended to be lower than predicted, and at^ery high strains where they
were larger than predicted. The effects at large strains are caused by fail
ure of the Gaussian distribution, eq (1.1.6) to apply for large extensions of
the chain molecules. This can be explained by non-Gaussian statistics [72].
RUBBER STRUCTURES AND PROPERTIES 19

Deviations at moderate strains have been attributed to inadequacy of


the stored energy function using only a single constant. Krigbaum and
Roe [74] analyze such deviations from the kinetic theory in detail, summa
rize the evidence that there are, in fact, appreciable energy changes in rub
ber deformations, and describe more recent attempts to test and refine this
theory.
Correction terms have been introduced into the theory to account for
free ends of chains which are not tied into the network and for chain en
tanglements which act as effective crosslinks [72, 73, 75, 76]. Fillers and
crystallites seem to provide fixed attachment points for network chains
and thus simulate to some extent the effects of additional crosslinks.
Strain-energy representation of rubber elasticity. The preceding sections
have shown how thermoelastic phenomena, along with rubber and ther-
modynamic analysis, led to the kinetic or statistical theory of rubber elas
ticity. In turn, this leads to a description of rubber elasticity in terms of a
stored energy function using a single elastic constant, called G in eq
(1.1.13).
The stored-energy approach has been developed for rubber in an en
tirely phenomenological way, independent of any molecular theory, both
by Mooney [77] and by Rivlin [78]. In applying the general elasticity the
ory for large deformations of incompressible, isotropic elastic materials,
Mooney assumed a linear shear stress law consistent with kinetic theory.
Rivlin showed that it was unnecessary, for many purposes, to assume any
particular elastic law and that the elastic law could be determined experi
mentally through relations derived from the theory.
The strain energy per unit unstressed volume stored in a material is a
function of the general components of strain at any point [79]. The stored-
energy function is unaffected by coordinate transformations. The form of
this function is as characteristic of the material as the stress-strain relation,
with which, of course, it is closely connected. The nature of the stored en
ergy function for a pure homogeneous strain completely determines the
elastic properties of the material. The assumption of incompressibility
simplifies the function, and is justified for rubber in practical terms be
cause stresses required for changing the volume are so much larger than
those required to change the shape.
It follows from this general theory for large elastic deformations of rub
ber [72, 79-81] that the stored energy W is a symmetric function of the
three principal extension ratios λ1, λ2, and λ3, and can be expressed in
terms of the three following strain invariants:

7, - \\ + \l + \l (1.1.23)
I^K-Xl + Xl-M + Xl-K, . (1.1.24)
I3 = \l-\l-\l ^ (1.1.25)
Assuming incompressibility, I3 = 1, and λ3 can be eliminated from eqs
(1.1.23) and (1.1.24) so that Ff can be expressed in terms of two independ
ent variables I1 and I2, which in turn contain only λ1 and λ2. This means,
of course, that only two of the extension ratios can be varied independ
ently.
The most general form of this stored energy function can be written [72]

3X (1.1.26)
20 MECHANICS OF PNEUMATIC TIRES

(/, - 3) and (/, - 3) are used in eq (1.1.26) instead of I1 and I2 so that W


will be zero for zero strain.
The first term of the series, i = 1, j = 0, gives for W the form derived
from the kinetic theory, eq (1.1.13) which gives a reasonably good first ap
proximation to the rubber stress-strain relations.
Retention of the first two terms, /' = 1, j = 0 and i = 0, j = 1 gives
ff=C,(/1-3) + CI(/2-3) (1.1.27)
This was the form derived by Mooney which, having two constants, pro
vides better agreement with experimental data than eq (1.1.13).
From eq (1.1.27) for simple extension, the force / per unit initial cross
section is derived as

f-2 (1.1.28)

Equation (1.1.28) is known as the Mooney-Rivlin equation. For simple


shear,
/. = 2(C, + C2)o (1.1.29)
in which /„ is shear stress, σ shear strain and hence 2(C1 + C2) is the modu
lus of rigidity.
Equation (1.1.28) has been exhaustively tested and debated [72, 74, 79].
There is little question that it gives better agreement with experiments at
moderate elongations than does eq (1.1.22) but the physical origin or sig
nificance of C2 is obscure. Ciferri and Flory [82] showed that it decreased
in value when better techniques were used to reach equilibrium strains,
but it evidently does not vanish entirely and sometimes is quite appre
ciable [74].
Energy effects have been shown to make a substantial contribution to
C2, but there are probably also entropy effects related to the orienting
properties of particular network structures as affected by chain length dis
tribution and chain packing [74]. The exact nature of this is still nebulous.
It would be necessary to retain more terms in eq (1.1.26) in order to ac
count for the upward curvature in rubber stress-strain curves at high elon
gation. These occur because of limited chain extensibility.
A convenient way to test eq (1.1.28) experimentally is to plot //2 (λ -
1/λ2) against 1/λ. If the equation holds, a straight line will be obtained,
the slope of which is C2 while the intercept is C1.

1.1.5. Rubber Viscoelasticity


Viscoelasticity, or delayed response to stress change, of rubber com
pounds is especially important for tire performance because it is greatly
enhanced by reinforcing carbon blacks. Equilibrium or reversible stress-
strain relations from the kinetic theory become submerged in viscoelastic
effects for such compounds. Tire deflections occur so rapidly that equilib
rium conditions are not approached for the rubber deformations in tires.
In general, rubber viscoelasticity becomes evident in a dependence of
properties on rate of deformation, time-deformation history, and temper
ature dependence in excess of the relatively small kinetic theory effect, ac-
RUBBER STRUCTURES AND PROPERTIES 21

LOAD REMOVED

PERMANENT
SET

1
I
LO
J
/ CONSTANT TIME
f- LOAD APPLIED

FIGURE 1.1.9 Typical deformation in creep, recovery, and permanent set.

cording to which the modulus should be proportional to absolute temper


ature. In recent years it has been found that tensile and tear strength
mechanisms [83] and even abrasion processes for rubber compounds [84]
are dominated by viscoelastic effects.
Figure 1.1.9 gives a diagram illustrating the delayed response of rubber
in a creep test where a test specimen is loaded with a constant weight.
There is a rapid deformation at the start followed by slow, gradual ap
proach to an equilibrium deformation. When the weight is removed, there
is a rapid retraction followed by a slow recovery which may never be com
plete, so that there is permanent set as well.
Figure 1.1.10 illustrates compressive stress-relaxation curves for an SBR
tread-type compound, when a test specimen was subjected to constant
compression. Rate of stress decay, i.e., time rate of change of stress at con
stant strain, for this compound evidently went through a minimum at
about 78° F. Stress relaxation at moderate temperature is largely a phys
ical phenomenon, but at elevated temperatures it is usually associated
with chemical changes such as oxidation and degradation.
Molecular and model concepts of rubber viscoelasticity. The molecular
basis for rubber viscoelasticity lies in viscous forces acting on the segments
of a chain molecule as they move in response to an applied stress. Each
segment is essentially drawn through a very viscous medium consisting of
its neighbors, so that its motion is retarded and an appreciable time is re
quired to adjust to a stress. Semiquantitative calculations based on such a
molecular picture show that it is essentially correct [75].
The mathematical theory of linear viscoelasticity can be presented by
an analysis of ideal spring-dashpot models to represent, respectively, the
elastic and viscous components of the response of the material to stress.
These models have been widely described [75, 85-87]. For stress relaxa
tion, that is, decay of stress at constant strain, it is convenient for the
model to consist of a large number of Maxwell elements in parallel such as
MECHANICS OF PNEUMATIC TIRES

FIGURE 1.1.10. Compressive stress relaxation for an SBR tread compound under 30%
compression.
shown in figure 1.1.1 1 (a). For descriptions of creep Voigt-Kelvin elements
in series are advantageous as shown in figure 1.1.1 l(b). Although any rub
ber viscoelastic curve, such as a creep or stress relaxation curve, can be
matched by a suitable model with sufficient elements [88], this is entirely
formal since the model elements cannot be identified with specific features
of molecular or network structure. If the material does not truly display
linear viscoelasticity, such complicated models cannot be expected to
cover a wide range of behavior. For a Voigt-Kelvin element, stress on the
unit at any time is the sum of the force arising from spring elongation and
that from dashpot velocity, so that one may write

(1.1.30)

Response of a Voigt-Kelvin element to a constant load is [75]


D(t) = e/a = [1 - exp (-t/r)\. (1.1.31)
Using the element to represent a unit cube of material, e in eqs (1.1.30)
and (1.1.31) is the strain, dt/dt the strain rate, σ the stress, E is Young's
modulus and τ is the retardation time of strain. This latter quantity is de
fined as Tj/E, where TJ is the internal friction of the material, that is, the
stress per unit strain rate. D(t), or e/o, is called the creep compliance.
If there are a number of elements in a series model, as in figure
1.1.1 l(b), there will be a discrete spectrum of retardation times, and creep
compliance will be given by a summation of the contributions to e from
the individual elements.
Viscoelastic relations between creep, stress relaxation, and complex dy
namic modulus. When the mathematical summation described above is
carried to the limit for a continuous distribution of retardation times of
strain, the distribution function is known as the retardation spectrum of
the material. The creep compliance is obtained from this distribution as an
integral expression [88, 89]. Similarly, the relaxation modulus can be de
rived as an integral expression involving the distribution function of the
relaxation times of stress, conveniently derived through a generalized
RUBBER STRUCTURES AND PROPERTIES 23

Maxwell model, figure 1.1. 11 (a). The relaxation and retardation spectra
are alternative ways of specifying the viscoelastic behavior of a vulcan-
izate. They are related by the mathematical theory of linear viscoelasticity
[88, 90]. In theory, these spectra can be determined for a material which
conforms to linear viscoelasticity from suitable stress relaxation or creep
curves. This is usually done by approximation methods. Similarly, when
the models are subjected to a sinusoidal driving force, the response can be
calculated by properly formulated integrations of either of the distribution
functions. Thus, in principle, any of these experimental methods, suitably
applied, can be used to completely determine the viscoelastic response of a
material to forces. For instance, data from a relaxation text can be used to
calculate creep [91, 92].
Superposition principle. A basic assumption of linear viscoelasticity the
ory is that the material responds to stresses in the same way regardless of
its past stress history. The effect of a change in stress can be superimposed
on effects remaining from previous stresses. This principle permits the re
sponse to be calculated for variable applied forces [88, 90].
As with the kinetic theory, the linear theory of viscoelasticity provides a
useful framework of reference from which to describe rubber properties.
Unfortunately it is not truly descriptive of tire compounds which are filled
with reinforcing carbon black. Reasons for this are apparent in light of the
restrictions of the superposition principle. In the first place, aside from
small shear stresses, rubber has a nonlinear stress-strain relation. In addi
tion there is an irreversible deformation component, or permanent set,
which can be described in particular cases by model elements but which
cannot be well described for all deformations. Finally, in vulcanizates
filled with reinforcing carbon black there appear to be stress-induced
changes in internal structure associated with breaking or slipping of poly
mer bonds, polymer-filler bonds or filler-filler bonds. These tend to invali
date application of the superposition principle and hence the linear theory
of viscoelasticity.

1.
G, \ G2,

,LU ftlil

(a) (b)
FIGURE 1.1.11. Models Jor viscoelasticity.
la) Maxwttt tltmtMi in parallel, (b) Voigl-Kebin dementi in xrla.
24 MECHANICS OF PNEUMATIC TIRES

Time-temperature superposition principle. A very useful viscoelastic con


cept is that time and temperature are equivalent for describing viscoelastic
behavior [75, 86, 89]. Thus a creep curve observed for short times at a
given temperature is identical with one observed for longer times at a
lower temperature, except that the curves are shifted on a logarithmic time
axis. They can be superimposed once more by proper scale changes on this
axis. Similarly, portions of a creep curve or stress relaxation curve can be
observed at different temperature, and these curve segments can then be
shifted along the log-time axis to construct a composite curve or master
curve, applicable for a given temperature, extending over many decades of
time. Figures 1.1.12 and 1.1.13 from Bueche [75] illustrate this procedure
for a plot of creep compliance against log-time. The shift factor for a curve
segment is designated «„ log a, being the horizontal displacement neces
sary to allow it to join smoothly into the master curve. This is the factor by
which the time scale is altered due to the difference in temperature, and is,
of course, a function of temperature.
For exact work, there is also a small vertical shift, modulus values being
multiplied by Tapa/Tp (or compliance values by the reciprocal ratio) to
take account of the entropy effect of temperature on stress. T0, and ρ0 are
absolute temperature and density, respectively, for standard conditions or
for the master curve, and T and ρ apply for the curve segment which is to
be shifted [86, 89].
It has been found [89] that for all linear viscoelastic materials over a
limited temperature range horizontal shift factors are given by the empiri
cal Williams-Landel-Ferry (WLF) equation:
log aT • -17.44(7- Ts)
(1.1.32)

SMtfT CACH POINT OH TMH


42°C CUBVf 0T UNITS CAi_CULATtQT USlf*6T<42*C
CM

15°
H -21 °
X

UJ
U

O
O

UJ
DC
O

LOG TIME , SEC.


FIGURE l.l.ll Plot of creep compliance vs. log lime for a natural rubber vulcanizate at a
series of temperatures [75].
RUBBER STRUCTURES AND PROPERTIES

o
I 3

O
o

LU
K
u
-2 8 10 12
LOG TIME , SEC.

FIGURE 1.1.13 Master curve, constructed by horizontal shifting of curve segments offigure
1. 1. 12, giving creep compliance at —56 °C, from reference f75J.

T, is the glass transition temperature of the material. Equation (1.1 .32)


provides quite satisfactory shift factors in the range Tt < T< Ts + 120 for
all types of viscoelastic phenomena. Either Kelvin or Centigrade temper
atures may be used in it. The applicability of eq (1.1.32) is often used as a
criterion for whether a process, such as abrasion or tearing of rubber is vis
coelastic in nature.
For tire deformations, this time-temperature equivalence principle im
plies that response of the rubber at high speeds will not be correctly eval
uated with conventional laboratory stress-strain testing speeds at the same
temperature, and that tire temperature and speeds should be taken into
consideration. To estimate the order of magnitude of the effect, assuming
tire temperature to be 25 °C and T = -70°C, eq (1.1.32) gives log aT =
-1 1.3 for T = 25°C and for T = 0°C, log aT = -10.0. Thus lowering the
test temperature from 25 °C to 0°C is roughly equivalent to increasing test
ing speed by a factor of 10", or 20, say from 20 in/min. to 400 in/min.
In the same way, high frequency vibration response of the rubber at
25°C could be simulated by testing at 0°C with frequencies lowered by a
factor of about 20. All this assumes that eq (1.1.32) is valid for tire com
pounds, which in general is not strictly so.
Dynamic properties. When a sinusoidal force is applied to a viscoelastic
material, the viscous reaction causes a lag of strain behind stress, just as in
a creep test, as illustrated in figure 1.1.14.
This phase lag is indicative of mechanical energy loss in each cycle, this
energy appearing as equivalent heat generation in the rubber. The amount
of energy loss is given by the area of the hysteresis loop when stress in the
material is plotted against strain, as shown in figure 1.1.15.
The dynamic modulus of a material is defined as a complex number
whose real part is the ratio of the stress component in phase with the strain
26 MECHANICS OF PNEUMATIC TIRES

STRESS
"N
STRAIN

181 cut

FIGURE 1.1.14 Diagram to illustrate phase lag ofstrain behind stress when a sinusoidalforce
acts on a rubber specimen.

to the strain itself, while the imaginary part is the ratio of the stress com
ponent 90° out of phase with the strain to the strain itself. The latter com
ponent, of course, is responsible for the energy losses. These relations are
expressed by
G* = G + iG" (1.1.33)
in which G* is the total complex shear modulus. The real component G′ is
called the storage modulus or the dynamic modulus, and G" is the loss
modulus. The complex Young's modulus is defined similarly and, assum
ing incompressibility, it is three times the complex shear modulus. Tan 5,

E" =

AREA OF ELLIPSE = 7rE"e02

FIGURE 1 . 1 . 1 5. Hysteresis loop showing its relations to storage and loss moduli and the phase
angle, S.
RUBBER STRUCTURES AND PROPERTIES 27

the tangent of the angular phase lag of strain behind stress is given by

tan 8=^-- (1.1.34)


G
Tan δ is a basic parameter for expressing the energy losses relative to the
energy stored. Losses in various dynamic test methods, such as rebound
experiments or decay of free vibrations, can all be expressed in terms of
tan δ [93].
All of these dynamic properties such as G′, G", and tan δ, being vis-
coelastic in nature, show marked dependence on temperature and fre
quency. However, in a limited range of mechanical frequencies, from
about five or ten to several hundred Hz and at temperatures well above
the glass transition temperature Tf the frequency dependence is rather
flat. Storage and loss moduli decrease with increase in temperature, show
ing that viscoelastic effects dominate over the increase predicted by the ki
netic theory. Tan δ also decreases with increase in temperature. This is a
saving feature for tire compounds, since otherwise they would tend to run
hotter and hotter in service rather than to approach an equilibrium tem
perature for a given operating condition.
To give some idea of numerical values, tan δ for tread compounds will
usually be found in the range 0.1 to 0.2 while dynamic modulus E′ will
range from 1500 to 2500 psi for low amplitude vibrations of about 60 Hz
at room temperature.
Dynamic modulus for tread compounds exceeds the static modulus,
sometimes by a factor of two or more, depending on the vibration condi
tions. It should be noted that the apparent Young's modulus for rubber
determined by either dynamic or static tests in compression depends upon
the shape of the test specimen unless the bearing surfaces are well lubri
cated. Thus the tread "buttons" in a tread pattern will be somewhat stiffer
in compression on a road which is dry than when it is wet. Empirical
shape factors [53, 94] such as the ratio of load area to free area are used to
take this shape effect into account. The effect is somewhat larger for dy
namic than for static tests [95].
Dynamic properties of tire tread compounds are further complicated by
a marked dependence on amplitude [96], giving rise to nonlinear effects.
Except for extremely small amplitudes [97], dynamic modulus decreases
with increasing amplitude, slowly approaching an asymptotic value at
high amplitudes. This gives rise to a skewed resonance curve as shown in
figure 1.1.16, obtained in this case by varying the mass of the system while
holding excitation frequency constant. These effects are structural or thi-
xotropic in character. However, for a steady forced vibration amplitude an
equilibrium structure is soon reached so that there is a sinusoidal response
without harmonics.
Energy losses in tires. Mechanical hysteresis losses in the rubber com
pounds and in the fabric of a tire result in a drag component manifesting
itself in internal heat generation which is readily observed as an increase
in tire temperature above ambient. Internal heat generation is especially
severe for thick, heavy-duty tires. It limits speeds and loads at which such
tires can operate, and it delayed the use of synthetic rubber in compounds
for airplane, bus, truck, and off-the-road tires.
In comparing energy losses for rubber compounds it is necessary to
specify the deformation circumstances. Hysteresis loss // per cycle per unit
28 MECHANICS OF PNEUMATIC TIRES

2.0

2500 3500 4500


MASS, GRAMS

FIGURE 1.1.16. Curve A, resonance curve for a natural rubber unfilled vulcanizate, forced
compressive vibrations, 60 Hz. Curve B, resonance curve for a tread compound showing
distortion because of nonlinear response.
Constant amplitude driving force.

volume under sinusoidal displacement is [98]


H = wooCo sin 8. (1.1.35)
Here o0 and e,, are stress and strain amplitude, respectively, and sin δ =
E"/|E*|.
If rubber compounds are compared at the same strain amplitude, a0 =
eo • |£*| and
(1.1.36)
If they are compared at the same stress amplitude, €<, o0/|£*| and
#„ = (1.1.37)
In both cases, energy loss is proportional to loss modulus. At constant
stress cycle amplitude stiffer compounds have lower losses because the loss
per cycle is inversely proportional to the square of the complex modulus.
Deformation conditions in a tire are so complicated that it is difficult"to
consider them to be purely constant stress amplitude or constant strain
amplitude cycles. Amplitude of bending cycles in the tire tread and body
do not depend much on compound modulus and so are approximately
constant strain amplitude cycles. Compressive stress cycles of the tread it
self are more like cycles at constant stress amplitude.
f Detailed experimental analysis to determine contributions of individual
I tire components of a 9.00-20 truck tire to drag [98] showed that tread com
pression contrihuted 32 percent of the drag, tread bending 27 ^percent,
body rubber 12 percent and cord system 29 percent.
Trie coefficient of rolling resistance of a tire, defined as the drag force to
pull the free rolling tire divided by the vertical tire load, is related to
power loss by the equation

(1.1.38)
60 SL
RUBBER STRUCTURES AND PROPERTIES 29
where R is the dimensionless coefficient of rolling resistance, P is power
loss, ft-lb/min.; S is speed, ft/sec.; and L is tire load, Ib. In dynamometer
tests on 9.00-20 track tires [99], power loss for a tire made with SBR com
pounds was about L5 times as large as that for a comparable tire made
with natural rubber compounds. This is symptomatic ofThe problem
which is encountered with heat generation in using synthetic rubber for
heavy duty tires.

1.1.6. Reinforcement of Rubber With Carbon Black


Carbon black is unique in tire technology. The benefits which ensue
from dispersing it in rubber are described by the ad hoc term "reinforce
ment." The carbon black/rubber system is very complex. It has been stud
ied exhaustively [100, 101] and many factors have been shown to contrib
ute to its effectiveness in tires. These include:
a. Chemical aspects, such as effects on vulcanization reactions through
adsorption of vulcanizing ingredients or through functional chemical
groups or elements on the carbon black surface
b. physical or chemical adsorption of segments of the chain molecules
c. physical bonding between carbon black particles resulting in carbon
black chains and a reticulated structure
d. microstress fields at the particles.
The complex interaction of these chemical, physical chemical, and phys
ical effects has precluded any simple, definitive explanation of all of the
phenomena. The fine carbon black particles are dispersed by high shear
ing forces in the thermoplastic rubber, where they tend to form a sort of
loose, flocculated or reticulated structure because of their surface activity
and mutual attractions. This structure is then interlaced by a network of
rubber chain molecules crosslinked during vulcanization. The carbon
black particles are fine enough and have sufficient surface activity to influ
ence the rubber network structure profoundly, perhaps furnishing more or
less stable junction sites which act like additional crosslinks to produce ef
fects not observed with an ordinary, diluent type of filler [101, 102]. There
is also a theory that the particles act to distribute the rubber network
stresses [75]. ,One effect is to greatly increase the viscous forces on the
chain segmeais_Qflhe rubber .molecuics"'pT^T5l6v(rio accomodate t o
stresses.
Carbon black in rubber increases the modulus or hardness, as do all fill
ers. Howgyf, distinctive improvements are observed in strength proper
ties such astensile; strength, tear strength and above all, in roadwear. On
the other handJ, mechanical energy losses, hysteresis loss and viscoelastic
respoflSesToTorces are greatly augmented. It is now realized that these two
broad characterizations of the effects of carbon black on physical proper
ties are closely connected. All types of strength failures in rubber probably
originate at small flaws and proceed by essentially a tearing process [50].
Viscoelastic mechanisms have been clearly demonstrated in tearing and
abrasion of unfilled vulcanizates [103, 104], and are incorporated in cur
rent theories of rubber tensile strength [83]. Stress relaxation and creep re
duce stress concentrations at the tip of a growing flaw or cut. This appears
as increased strength. There is an additional strengthening mechanism with
carbon black in that enhanced stresses at the particles produce molecular
30 MECHANICS OF PNEUMATIC TIRES

TABLE 1.1.3. Description ofseveral carbon blacks [107]


Av. panicle Surface area. Apparent sp.
diam. A (Elec m2/g (BET N2 vol. at 734 psi,
Type tron microscope) adsorption) cc/g
HMF (High Modulus Furnace 460-660 30-45 1 20-1 46
^FEF (Fine Extrustion Furnace) 310-580 36-48 1.40-1.70
HAF (High Abrasion Furnace) 260-350 62-88 1.45-1.60
ISAF (Intermediate Super
Abrasion Furnace) 175-275 95-135 1.45-1.75
SAF (Super Abrasion Furnace) 140-270 120-145 1.55-1.75

orientation or alignment, a sort of fibering, which blunts the tear tip and
tends to divert the tear from a straight line course.
f Carbon blacks are characterized by particle size, surfacearea, and struc
ture. Particle size is measured from electron micrograprls. "Surface area is
•• determined by iodine [105] or nitrogen adsorption. "Structure" measures
the proclivity of a carbon black to form reticulated structures. It is eval
uated by oil absorption tests [106] or measurement of packing volume un
der pressure. Manufacturing methods for carbon blacks are so advanced
that they can now be produced commercially with practically any desired
combination of these three characteristics. Table 1.1.3 summarizes ranges
of these properties for carbon blacks used frequently in tires.

4000 •

"0 25 50 75 100
CARBON BLACK LOADING, PTS./IOO PTS. POLYMER

FIGURE 1.1.17. Effect of loadings of ISAF carbon black in SBR-1500C on tensile strength
and pendulum rebound.
RUBBER STRUCTURES AND PROPERTIES 31

800

0 28 SO 75 100
CARBON BLACK LOADING, PTS./IOO PTS. POLYMER

FIGURE 1.1.18. Effect of loadings of ISAF carbon black in SBR-1500C on breaking


elongation and hardness.

Figures 1.1.17 and 1.1.18 display the effect of carbon black loadings on
tensile strength, pendulum rebound, breaking elongation, and hardness
[108].
Tensile strength usually goes through a broad maximum as carbon
black loading increases, as does abrasion resistance. The level of 45-55,
parts of black per hundred of rubber in tire treads, if not the optimum for
wear in a particular compound, will generally represent the best balance
between wear, resistance to tread cracking, heat generation, traction, etc.,
and give best overall performance.
Stress-strain curves for carbon black filled vulcanizates show pro
nounced stress softening [109, 110], that is, the rubber compound has a
much lower modulus on the second extension than on the first, as shown
in figure 1.1.19. This effect was widely studied and for a long time was
thought to be an important characteristic of the reinforcement process. It
is now known to be a much more general phenomenon [111, 112] and its
significance for reinforcement has been obscured. This effect is obviously
related to the fact that stable reproducible mechanical properties for tires
are observed only after a "break-in" run.
Abrasion of rubbeLOBVolves very complicated failure processes [50,
103], including softening and fatigue of a thin surface layer, probably as-
32 MECHANICS OF PNEUMATIC TIRES

100 200
ELONGATION. %

FIGURE 1.1.19 Stress-softening for a tread vulcanizate illustrating approach to an


equilibrium hysteresis loop after many cycles.

140

10 20 SO 40 60 60 100 ISO 200

SURFACE AREA , m2/fl

FIGURE t.1.20 Correlation of treadwear with specific surface area of the carbon black.
Data assembled by Sludebaker [1 13] from icveral aourca.
RUBBER STRUCTURES AND PROPERTIES 33

RELATIVE WEAR RESISTANCE


I20n

HI6H
STRUCTURE
BLACK

ISAF SMALL
PARTICLE
ISAF-HS SIZE
NORMAL
STRUCTURE * HAF-HS
BLACK LARGER
PARTICLE
SIZE
• HAF

I VERY
H- SEVERE -*4-» ••^-MODERATE-

0 20 40 60 80 100
TEST SEVERITY, MILES/MIL OF WEAR

FIGURE 1.1.21 Relations between relative treadwear, particle size and "structure" of carbon
black and severity of service (113].

sociated also with oxidative deterioration and actual smearing under se


vere conditions, localized cutting and chipping from road asperities, and
shearing off and rolling up of thin flakes of rubber from the surface. Car
bon blacks have evolved largely from requirements to improve treadwear,
and have been developed through tire experience and carefully controlled
road tests. No one type of black can be optimum for the wide range of
rubber compounds, tires, and service conditions. A broad, general correla
tion exists between specific surface area of the carbon black and treadwear
[113] as shown in figure 1.1.20. This emphasizes the importance of surface
interactions between rubber and filler for good wear resistance. Figure
1.1.21, however, reflects experience showing that high structure in carbon
black is especially advantageous for very severe service [113]. This is an
indication of the present sophistication in development and use of carbon
blacks in tire compounds.

References
[1] Walker, Richard, ed., Materials and Compounding Ingredients for Rubber and Plas
tics (Rubber World, New York, N.Y. 1965).
[2] Brown, R. J., et •!.. Rubber World 145(2), 70 (1961); Rubber Chem. Tech. 35(2), 546
(1962).
[3] Sarbach, D. V., Hallman, R. W , and Cavicchia, M. A., Rubber Age 98(1 1), 67 (1966).
34 MECHANICS OF PNEUMATIC TIRES

[4] 1967 Annual Book of AS I M Standards, Pan 28, Rubber, carbon black, gaskets,
Method D2226-65T, Description of types of petroleum extender oils, p. 992 (Am.
Soc. Testing Mats., Phil., Penn.).
5] Storey, E. B., Rubber Chem. Tech. 34(5), 1402 (1962).
6] Stout, W. J., and Eaton, R. L., Rubber Age 99(12), 82 (1967).
~ Ref. [4], p. 844.
8] Snydcr, J. W., and Leonard, M. II Carbon black, chap. 8, Introduction to Rubber
Technology, Morton, M., ed., p. 172 (Reinhold Publishing Corp., New York, N.Y.,
1959).
[9] Studebaker, M. L., Rubber Chem. Tech. 30(5), 1400 (1957).
[10] Alliger, G., and Sjothun, I. J., eds., Vulcanization of Elastomers (Reinhold Publishing
Corp., New York, N.Y., 1964).
[11] Hofmann, W., Vulcanization and Vulcanizing Agents (Palmerton Publishing Co.,
New York, N.Y., 1967).
[12] Bateman, L., et al., Chemistry of vulcanization, chap. 15, The Chemistry and Physics
of Rubber-Like Substances, L. Bateman, ed., p. 449 (John Wiley & Sons, New York,
N.Y., 1963).
[13] Trivette, Jr., C. D., Morita, E., and Young, E. J., Rubber Chem. Tech. 35(5), 1360
(1962).
[14] Ambelang, J. C., et al., Rubber Chem. Tech. 36(5), 1497 (1963).
[15] Thornley, E. R., Trans. Inst. Rubber Ind. 40, T1 (1964); Rubber Chera. Tech. 37, 973
(1964).
[16] Rubber: Supply and distribution for the United States, Current Industrial Reports,
U.S. Dept. of Commerce, Bureau of the Census, Industry Div., Washington, D.C.,
Dec. 1967.
[17] Description of Synthetic Rubber and Latices (International Institute of Synthetic Rub
ber Producers, Inc., New York, N.Y., Jan. edition 1968); Ref. [4], p. 1150.
18 Buckley, D. J., Rubber Chem. Tech. 32(5), 1475 (1959).
19 Dudley, R. H., and Wallace, A. J., Rubber World 152(2), 66 (1965).
20 Andrews, J., Rubber Age 99(12), 53 (1967).
21 Hall, W. S., and Norman, D. T., Rubber Age 93(1), 77 (1963).
22 Vanderbilt News 28(4), 47 (1962): for other tire compound formulations see 31(1), 11
(1965); 31(2), 8 (1965); 32(2), 7 (1967).
[23] K IK hunt, C. J., Changing trends in tread type carbon blacks. Paper presented at meet
ing of the ACS Division of Rubber Chemistry, Cleveland, Ohio, April 23-26, 1968;
Abstract in Rubber Age 100(3), 79 (1968).
[24] Vance, R. M., and Burgess, K. A., Laboratory tire groove cracking test, Paper pre
sented at meeting of ACS Divison of Rubber Chemistry, Cleveland, Ohio, April 23-
26, 1968; Rubber Chem. Technol. 41(4), 1080 (1968).
[25] Ref. [4], Method D15-66T, Sample preparation for physical testing of rubber products,
p. 1.
26] Juve, A. E., Physical testing, chap. 19, p. 462, in Ref. [8].
27] Scott, J. R., Physical Testing of Rubbers (Palmerton Publishing Co., New York, N.Y.,
1965).
[28] International Organization for Standardization Recommendations (Listed in ISO
Catalog ( 1967), USA Standards Institute, New York); also listed in Ref. [4], p. 1 121.
[29] Franck, A., Hafner, K., and Kern, W. F., Kaut. Gummi Kunstst 13(12) WT 292
(1960); Rubber Chem. Tech. 35(1), 76 (1962).
[30] Gehman, S. D., and Ogilby, S. R. ASTM STP-383, Continuous Measurement of the
Cure Rate of Rubber (Am. Soc. Testing Mats., Phil., Penn. 1965).
[31] Ref. (4], Method 1)4 12 66, Tensile testing of vulcanized rubber, p. 200.
[32] Gehman, S. D., Chem. Revs. 26(2), 203 (1940).
33] Mandelkern, L., Crystallization of Polymers (McGraw-Hill Book Co., New York,
N.Y., 1964).
[34] Mullins, L., and Tobin, N. R., Trans. Inst. Rubber Ind. 33(1), 2 (1957); Rubber Chem.
Tech. 31, 505 (1958).
[35] ASTM tentative methods D2704, D2705, and D2706 on methods of measurement of
curing characteristics. Should appear in future editions of Ref. [4].
[36] Gehman, S. D., Maxey, F. S., and Ogilby, S. R., Rubber Chem. Tech. 38(4), 757
(1965).
[37 Gehman, S. D., Rubber Chem. Tech. 40(1), 36 (1967).
[38 Freeman, H. A., Rubber Age 90(5), 779 (1962).
39 Ref. [4], Methods D454-53, D572-67, D573-53, D865-62.
Barnhart, R. R., and Newby, T. H., Antioxidants and antiozonants, chap. 6, p. 130, in
Ref. (8).
[41] Buist, J. M., Aging and Weathering of Rubber (W. Heffer & Sons, Cambridge, Eng
land, 1956).
RUBBER STRUCTURES AND PROPERTIES 35

[42] Shelton, J. R., Rubber Chem. Tech. 30(5), 1251 (1957).


[43] Norling, P. M., Lee, T. C. P., and ToboUky, A. V., Rubber Chem. Tech. 38(5), 1 198
(1965).
(44) Ref. [4], Method D518-61, Resistance to surface cracking of stretched rubber com
pounds, p. 275.
[45] Ref. [4], Method Dl 149-64, Accelerated ozone cracking of vulcanized rubber, p. 554.
[46] Symposium on Effect of Ozone on Rubber, ASTM STP-229 (Am. Soc. Testing Mats.,
Phil., Penn., 1958).
[47] Andrews, E. H., et al, Ozone attack on rubbers, chap. 12, p. 329, in Ref. [12].
[48] Ref. [4], Method D624-54, Tear resistance of vulcanized rubber, p. 342.
[49] Kainradl, P., and Handler, F., Kautschuk Gummi 12, WT 239 (1959); Rubber Chem.
Tech. 33, 1438 (1960).
[50] Gehman, S. D., Mechanism of tearing and abrasion of reinforced elastomers, chap. 2,
Reinforcement of Elastomers, G. Kraus, ed., p. 23 (Interscience Publishers, Inc.,
New York, N.Y., 1965).
[51] Ref. [4], Method D2240-64-T, Indentation hardness of rubber and plastics by means of
a durometer, p. 1009.
[52] Ref. [4], Method D1415-62T, International hardness of vulcanized natural and syn
thetic rubbers, p. 642.
[53] Payne, A. R., and Scott, J. R., Engineering Design with Rubber, p. 122 (Interscience
Publishers, New York, N.Y., 1960).
[54] Waters, N. E., Brit. J. Appl. Phys. 16, 557 & 1387 (1965); J. Inst. Rubber Ind. 1(1), 51
(1967).
[55] Coddington, D. M., Marsh, W. D., and Hodges, H. C., Rubber Chem. Tech. 38(4), 741
(1965).
[56] Ref. [4], Method D 1054-66, Impact resilience and penetration of rubber by the
rebound pendulum, p. 509.
[57] Ref. [4], Method D945-59, Mechanical properties of elastomeric vulcanizates under
compressive or shear strains by the mechanical oscillograph.
[58] Gehman, S. D., Rubber Chem. Tech. 30(5), 1202 (1957).
[59] Ref. [4], Method D2231-66, Forced vibration testing of vulcanizates, p. 1003.
[60] Harris, C. M., and Crede, C. E., eds.. Shock and Vibration Handbook, Vol. 3, pp. 36-
25 to 36-27 (McGraw-Hill Book Co., New York, N.Y., 1961).
[61] Ref. [4], Method D623-62, Compression fatigue of vulcanized rubber, p. 334.
62] Ref. [4], Method D430-59, Dynamic testing for ply separation and cracking of rubber
products, p. 227; Method D813-59, Resistance of vulcanized rubber or synthetic
elastomers to crack growth, p. 410.
[63] Beatty, J. R., Rubber Chem. Tech. 37(5), 1341 (1964).
[64] Ref. [4], Method D394-59, Abrasion resistance of rubber compounds, p. 190; Method
D2228-63T, Abrasion resistance of rubber and elastomeric materials by the Pico
method, p. 985.
[65] Vogt, W. W., Ind. Eng. Chem. 20, 303 (1928).
[66] Lambourn, L. J., Trans. Inst. Rubber Ind. 4, 210 (1928).
[67] Davison, S., et al., Rubber World 151(5), 81 (1965); 151(6), 79 (1965); Rubber Chem.
Tech. 38(3), 457 (1965).
[68] Meyer, K. H., and Fern, C., Helv. Chirn. Acta 18, 570 (1935).
[69] Anthony, R. L., Caston, R. H., and Guth, E., J. Phys. Chem. 46, 826 (1942).
[70] Wood, L. A., and Roth, F. L., J. Appl. Phys. 15, 781 (1944).
[71] Gee, G., Trans. Faraday Soc. 42, 585 (1946).
[72] Treloar, L. R. G., The Physics of Rubber Elasticity, second edition (Oxford University
Press, 1958).
[73] Flory, P. J., Principles of Polymer Chemistry (Cornell University Press, Ithaca, New
York, 1953).
[74] Krigbaum, W. R., and Roe, R. J., Rubber Chem. Tech. 38, 1039 (1965).
[75] Bueche, F., Physical Properties of Polymers (Interscience Publishers, New York, N.Y.,
1962).
[76] Mulhns. L., and Thomas, A. G., Theory of rubber-like elasticity, chap. 7, p. 155, in
Ref. [12].
[77] Mooney, M., J. Appl. Phys. 11, 582 (1940).
[78] Rivlin R. S., chap. 10, vol. 1, Rheology: Theory and Application (5 vols.), F. R. Eir-
ich, ed., p. 351 (Academic Press, New York, N.Y., 1956).
[79] Varga, O. H., Stress-Strain Behavior of Elastic Materials; Selected Problems of Large
Deformations (Interscience Publishers, New York, N.Y., 1966).
[80] Rivlin, R. S., Phil. Trans. Roy. Soc. (London) A241, 379 (1948).
[81] Blatz, P. J., Rubber Chem. Tech. 36(5), 1450 (1963).
[82] Ciferri, A., and Flory, P. J., J. Appl. Phys. 30, 1498 (1959).
[83] Halpin, J. C., Rubber Chem. Tech. 38(5), 1007 (1965). '
36 MECHANICS OF PNEUMATIC TIRES

[84] Schallamach, A., Rubber Chem. Tech. 41(1), 209 (1968).


[85] Alfrey, Jr., T., Mechanical Behavior of High Polymers, Vol. 6 of High Polymers (Inter-
science Publishers, New York, N.Y., 1948).
[86] Tobolsky, A. V., Properties and Structure of Polymers (John Wiley & Sons, New
York, N.Y., 1960).
[87] Flilgge, W., Viscoelasticity, (Blaisdell Publishing Co., Waltham, Mass., 1967).
[88] Leaderman, H., chap. 1, vol. 2 of Rheology: Theory and Application (5 vols), F. R.
Eirich, ed., p. 1 (Academic Press, New York, N.Y., 1958).
[89] Ferry, J. D., Viscoelastic Properties of Polymers (John Wiley & Sons, New York, N.Y.,
1961).
[90] Gross, B., Mathematical Structure of the Theories of Viscoelasticity (Hermann & Cie,
Paris, France, 1953); J. Appl. Phys. 18, 212 (1947); 19, 257 (1948); 22, 1035 (1951).
Hopkins, I. L., and Hamming, R. W., J. Appl. Phys. 28(8), 906 (1957).
Aklonis, J. J., and Tobolsky, A. V., J. Appl. Phys. 36(11), 3483 (1965).
Gehman, S. D., Rubber Chem. Tech. 30(5), 1202 (1957).
Kimmich, E. G., India Rubber World 103(3), 45 (1940).
Gehman, S. D., J. Appl. Phys. 11(6), 402 (1943).
Gui, K. E., Wilkinson, Jr., C. S., and Gehman, S. D., Ind. Eng. Chem. 44, 720 (1952).
Payne, A. R., Dynamic properties of filler-loaded rubbers, chap. 3, p. 69, in Ref. [50].
Collins, J. M., Jackson, W. L., and Oubridge, P. S., Trans. Inst. Rubber Ind. 40, 239
(1964); Rubber Chem. Tech. 38(2), 400 (1965).
[99] Stiehler, R. D., et al., Proc. Intern. Rubber Conf., p. 73, Washington, D.C., 1959.
[100] Kraus, G., ed., Reinforcement of Elastomers (Interscience Publishers, Inc., New York,
N.Y., 1965); Rubber Chem. Tech. 38(5), 1070 (1965).
[101] Parkinson, D., Reinforcement of Rubbers, Monograph of the Inst. Rubber Ind. (Lon
don, 1957).
[102] Bueche, A. M., J. Polymer Sci. 25, 139 (1957).
[103] Schallamach, A., Rubber Chem. Tech. 41(1), 209 (1968).
[104] Greensmith, H. W., Mullins, L., and Thomas, A. G., Strength of rubbers, chap. 10, p.
249, in Ref. [12].
[105] Ref. [4], Method D1510-65, Iodine adsorption number of carbon black, p. 710.
[106] Ref. [4], Method D2414, Dibutyl phthalate absorption number of carbon black, p.
1044.
[107] Studebaker, M. L., Compounding with carbon black, chap. 12, p. 319, in Ref. [50].
[108] Data from Goodyear Tech Book for the Chemical Process Industry, The Goodyear
Tire & Rubber Co., Chemical Div., Akron, Ohio 44316.
[109] Dannenberg, E. M., and Brennan, J. J., Rubber Chem. Tech. 39(3), 597 (1966).
[110] Kraus, G., Childers, C. W., and Rollmann, K. W., J. Appl. Polymer Sci. 10, 229
(1966); Rubber Chem. Tech. 39(5), 1530 (1966).
[111] Trick, G. S., J. Appl. Polymer Sci. 3(8), 252 (1960).
[112] Harwood, J. A. C., Mullins, L., and Payne, A. R., J. Appl. Polymer Sci. 9, 301 1 (1965);
Rubber Chem. Tech. 39(4), 814 (1966).
[113] Studebaker, M. L., Rubber Chem. Tech. 41(2), 373 (1968).
Chapter 2
TIRE CORD AND CORD TO
RUBBER BONDING
T. Takeyama,1 J. Matsui,1 M. Hijiri1

2.1. Physical properties of tire cords 38


2. 1 . 1 . Introduction 38
2.1.2. High tenacity rayons 39
2.1.3. High tenacity nylons 43
2.1.4. Polyester cord 55
2.1.5. Steel wire cord 60
2.1.6. Fiber glass cord 61
2.1.7. Wholly aromatic polyamides 63
2.1.8. Miscellaneous cords 63
2. 1 .9. Cordless tires 65
2.1.10. Comparative analysis of various tire cords 65
2.1.11. Impact resistance of tire cords 68
2.1.12. Fatigue resistance of tire cords 73
2.2.13. Effects of twist on cord properties 80
2.2. Rubber-to-cord bonding <95
2.2.1. Introduction 85
2.2.2. Outline of bonding methods 86
2.2.3. Adhesion mechanism 87
2.2.4. Adhesive treatment of nylon and rayon-RFL treat
ment 92
2.2.5. Adhesive treatment of polyester 96
2.2.6. Adhesive treatment of miscellaneous tire cords 106
2.2.7. Evaluation of adhesion 107
References 115
1 Tony Industries, Inc., Tokyo. Japan

37
38 MECHANICS OF PNEUMATIC TIRES

2.1. Physical Properties of Tire Cords


2.1.1. Introduction
The first pneumatic tire was made in 1888 by J. B. Dunlop with Irish
flax as thereinforcing material. This fiber was one of the strongest aTthat
time-TTwaTgradually replaced by cotton, since Irish flax was expensive.
Cotton tire fabrics in the early stages, about 1910, were plain fabrics. As
requirements imposed by the severity of tire operating conditions in
creased, these fabrics were gradually replaced by the present tire cord fab
rics, which were first devised by J. F. Palmer in 1892.
Prior to World War II, cotton was the sole textile used to any large ex
tent in pneumatic tires. However, this cotton tire cord fabric also failed to
meet requirements imposed by increasing severity of tire operations, so
that tire engineers began to consider man-made fibers.
The first rayon tire cord tenacity was about two grams per denier, and
was produced in 1923. Du Pont started to manufacture high tenacity
rayon, Cordura, in 1933 [1]2 and Courtaulds also started to manufacture
high tenacity rayon, Tenasco, in 1936 [1], but by 1940 high tenacity rayons
had only a small portion of the total tire cord market. In the late 1940's,
use of high tenacity rayon tire cords increased rapidly in the United States
and Europe. In Japan, rayon truck tires were first manufactured in 1951
[2-3]. Initially rayon cords were used only for truck tires. The rayon cord
tire had improved carcass performance and its life was increased 30 to 60
percent [2-3]. This improvement of tire quality was utilized more for truck
tires than for passenger car tires. Increased power of automobile vehicles,
however, gave birth to troubles in rayon cord (truck) tires. New, tougher
materials were required for tire cords, especially for heavy duty tires.
In 1947, nylon cords (nylon 66) were examined as reinforcing materials
for truck tires in the United States, and it was confirmed that nylon truck
tires have excellent properties, especially under severe operating condi
tions. In the late 1950's, rayon cords were gradually replaced by nylon
cords, especially in the United States.
Japanese tire manufacturers made extensive efforts to use nylon 6 for
tire cords. Nylon 6 was available from domestic suppliers (nylon 66 was
more costly) and nylon 6 tires, when post cure inflation was used, showed
no difference from nylon 66 tires in practice [2], [3]. Mass production of
nylon tires in Japan started in 1958. In the first five years, 60 percent of all
tire cords had been replaced by nylon cords. In 1967 nylon cords were
used for 90 percent of the tires made in Japan.
In Europe, rayon cord still had a large portion of the tire cord market in
the late 1960's.
Polyester tire cord does not exhibit flatspotting and is peculiarly suitable
for radial and belted-bias tires. The price oT this cord has become lower
than that oT nylon cord.
Goodyear started to produce polyester tires for passenger car service in
1962. By 1969, production of polyester tires had increased rapidly in the
United States, and polyester cord had become very important in the field
of passenger car tires.
Mass production of polyester tires in Japan started in 1967 and their
production is gradually increasing. In 1975, 50 percent of bias tires and 30
percent of radial tires for passenger cars were made from polyester cord.
In spite of the market assault by nylon and polyester, rayon cold is one
1 Figures in bracked indicate the literature references at the end of this chapter.
TIRE CORD AND CORD TO RUBBER BONDING 39

of the most suitable materials for the belt cords of radial and belted-bias
tires, and still keeps its position in the original equipment market for pas
senger car tires. However, the increase in production cost of rayon has
gradually depressed the production output, especially in the United States
and Japan.
Steel cord, which made its appearance in France in 1936, was used ini
tially in Europe and now has become one of the important tire cords in the
United States and Japan as well. The steel share of the European tire cord
market is 54 percent for truck and bus and 35 percent for passenger car.
Estimates for 1980 are 71 percent and 51 percent, respectively.
Glass cord, which appeared as a belt cord for bias-belted tires in 1966,
has been used mainly in the United States.
Wholly aromatic polyamides were initially used as tire cords in 1970.
They are characterized by their extremely high modulus and tenacity.
Kovac has summarized the history of tire cord in figure 2.1 [4].
2.1.2. High Tenacity Rayons
For more than thirty years, high tenacity rayons have been used as tire
cords. The tenacity of the first rayon tire cord in 1923 was only two grams
per denier as previously mentioned. In the 1930's, commercial production
of high tenacity rayons, Cordura and Tenasco, started in the United States
and in England, respectively. Tenacity of these cords was about 2.3 grams
per denier.
Thereafter, tenacity and durabi lity of rayon cords was improved and su
per series rayons, (e.g. Super I, Super II and Super III) were developed.
More recently, newer types of rayon cords, compressed rayon Dynacor [5-
7], and extra high modulus rayon were developed [8].
The improvement of cord tenacity of rayons is illustrated in figure 2.2
[9]. Cord tenacity of rayon now has reached five grams per denier or
higher.
The production process for high tenacity rayons differs in several re-

400-

1940 1950 1960 1970 1980


FIGURE 2. 1 United Slates tire cord usage [4J.
MECHANICS OF PNEUMATIC TIRES

1950 1955 I960 1965 1970

FIGURE 2.2 Evolution of rayon lire cord strength.

spects from that for regular rayons. High tenacity rayons have to be made
denser and more uniform in fine structure than regular rayons.
High quality wood pulp, and high quality, homogeneous viscose solu
tion are usually used in the production of high tenacity rayons.
The degree of polymerization of cellulose in the production of high
tenacity rayons ranges from 400 to 600, which is considerably greater than
that of regular rayon, 300. A higher concentration of carbon disulfide and
sodium hydroxide, and lower concentration of cellulose, are employed to
improve the solubility of viscose.
The high yarn strength can be obtained upon application of stretch at
high temperature.
Cox [10] first found that some additives (retardants) in the spinning
bath were effective for this purpose. These modifications by retardants be
came very important in the production of super series rayons. These retar
dants have the effect of retarding the regeneration of the cellulose from the
xanthate and of increasing the tenacity of the fiber [1]. A filament cross
section obtained by this method is circular and more uniform, as demon
strated by absence of skin and core regions, which can be seen in cross sec
tions for earlier high tenacity rayons. Many reports were published on the
effects of retardants on regeneration and coagulation mechanisms. It was
explained that these retardants react with zinc ions to form stable colloidal
chelate compounds, suppress diffusion of the coagulant and then slow
down the rate of regeneration.
Yarn-to-cord tenacity conversion efficiency of rayon cords is quite low.
Therefore, choice of cord lubricant is important. A lubricant, which de
creases the friction coefficient of the yam and improves the yarn-to-cord
tenacity conversion, has to be developed [11-12]; see table 2.1 [11].
Detailed investigations of fine structure of super high tenacity rayon fil
aments show that while overall change in total crystaUinity is very small,
the average crystallite size has been reduced. Thus, a fine, even textured
filament permits the load to be distributed more evenly across its struc
tural elements.
Internal structure of a rayon filament is simply expressed in terms of lat
eral order distribution [13-16] which can be measured by accessibility of
the cellulose internal structure to various chemicals, e.g., esterification by
formic acid, dissolution by alkali solutions and so on. Skin or less-ordered
TIRE CORD AND CORD TO RUBBER BONDING 41

TABLE 2.1. Correlation of cord strength with yam fric


tion coefficient [11],
Yarn friction Bone dry
Specimen coefficient strength
(1650/2 per 100
rayon cords) Static Dynamic denier
A 0.207 0.197 263
B .200 .184 266
C .160 .150 272
D .132 .158 280
E .107 .144 296
F .118 .133 306

regions can be dissolved by alkaline solution more easily than core or or


dered regions.
The change of lateral order distribution in super series rayon filaments
is illustrated in figure 2.3 [17]. The improvement of super series rayons ex
plained by change of the lateral order distribution to a low ordered, nar
row distribution.
Dynacor rayon (compressed rayon [3] [5]): Dynacor rayon, manufactured
by all Ty rex members, has the same chemical composition as conventional
rayon tire cord, but differs in physical characteristics and in the way it is
processed. It is made of low modulus, high tenacity and twisted greige
yam dipped at minimum tension in a standard RFL (Resorcinol Formal
dehyde Latex) adhesive and immediately stretched 15 percent prior to
drying.
This processing technique increases dip penetration into the cord.
Stretching aligns the filaments and compacts the bundle. Physical locking-
in of the dip and uniform stress distribution among the several thousand
individual filaments in the cord are obtained. In addition, it is said that
longitudinal air wicking, considered by many tire experts to be a major
cause of ply separation, is reduced to a minimum and cord-to-rubber
bonding made more durable. Equally important, this cord is claimed to
cost no more than regular rayon and requires little or no modification of
processing equipment.
Performance results together with indoor wheel tests [3] are claimed to
indicate that this rayon cord has greatly improved resistance to ply separa
tion, and that in the accelerated test for tread separation, Dynacor tires
SUPER A

ORDER
FIGURE 2.3 Change of lateral order distribution of super series rayons.
42 MECHANICS OF PNEUMATIC TIRES

run three to four times longer than regular rayons, with separations gener
ally occurring at a rubber to rubber interface.
Extra high modulus rayon [6]: Extra high modulus rayon is a polynosic
type of filament yarn. Extra high modulus rayon yarn has a tenacity of
more than 9 g. per denier bone dry and approximately 8.5 g. per denier in
the conditioned state. Elongation at break, 4 to 5 percent bone dry and 5
to 6 percent conditioned, is extremely low, while wet strength reaches 7 g.
per denier and wet elongation at break only 6 percent.
Young's modulus is very high, (see figure 2.4 [8]), and its temperature
dependence is low. This indicates dimensional stability for extra high
modulus rayon in tire construction, particularly for radial ply tires. Effects
of twist on cord strength of Super III rayon and extra high modulus rayon
are indicated in figure 2.5a [8]. The curve for Super III indicates increase
in strength to 22 kg. in the twist range of 20 to 40 turns per 10 cm., fol
lowed by a downward trend because of increasing conversion losses. With
low-extension, extra high modulus rayon cord, the curve starts out at 26
kg. and then shows a steady downward trend over the entire range repre
sented to reach 15 kg. In one range of twist extra high modulus rayon cord
has a greater breaking load than Super III, while in the higher range the
breaking load of Super III is superior.
Results of the Firestone flex fatigue test are illustrated in figure 2.5b [8].
The shape of the curve shows that higher cord twist can increase fatigue
resistance. A twist level of approximately 47 turns per 10 cm. gives an op
timum balance between pre- and post-flexing strength. The curve for extra
high modulus rayon cord shows great similarity to that for Super HI, ex
cept that retained strength at all twist levels is 5 kg. lower. In the com-
STEEL
EXTRA HIGH MODULUS RAYON
SUPER IE RAYON

3 6 5 6
ELONGATION ( %, )
FIGURE 2.4 Comparison ofstress-strain curves ofextra high modulus rayon and other cords.
TIRE CORD AND CORD TO RUBBER BONDING

20 30 7,0 50 60
TWIST (TURNS/10CM)
FIGURE 2.5 Effect of twist on cord tensile strength andfatigue resistance of Super III rayon
and extra high modulus rayon [8].
(a) Tensile strength vs. twin.
(b) Residual tensile strength after flexing with Firestone Flex Tester vs. twist

mercial twist range, breaking loads both before and after the flex test are
lower with extra high modulus rayon than with Super III. Use of this low-
extension rayon cord for conventional tire construction seems out of the
question. On the other hand, its high strength and low elongation may
make extra high modulus rayon cord a choice for reinforcing belts in
radial ply tires, where it is not subjected to such severe flex compression.

2.13. High Tenacity Nylons


Nylon cord is the most popular tire cord. In the standard twist construc
tion 840D/2, 47 × 47 (turns per 10 cm.), the cord strength of current nylon
6 and nylon 66 are 8 g. per denier and 7.5 g. per denier, respectively.
As compared with rayon cords, nylon cords give much longer tire life,
particularly when used in heavy duty truck tires. For this reason, nylon
cord was initially applied in heavy duty truck tires and then was gradually
adopted for light truck tires and replacement tires for passenger cars. Al
most all nylon cords used in Japan are nylon 6. On the other hand, nylon
66 cords have a greater portion of the nylon cord market in the United
States. Nylon 6 cord differs slightly from nylon 66 cord in regard to tire
cord characteristics. Nylon 6 cord has greater tenacity, is more economical
and more shock resistant than nylon 66 cord. Nylon 6 cord balances so
well with carcass compounds that dynamic adhesion fatigue is highly im
proved [2]. Furthermore, Allied Chemical stated their nylon 6 tire cord,
Caprolan, has better heat aging resistance, adhesion to rubber, and flex fa
tigue resistance than nylon 66 [1].
44 MECHANICS OF PNEUMATIC TIRES

On the other hand, nylon 66 has better thermal stability than nylon 6
due to polymer melting point difference. Thermal shrinkage at fairly high
temperatures such as tire curing temperatures is greater than for nylon 66;
see figure 2.6 [18]. This requires minor changes in processing conditions in
manufacturing tires with nylon 6 cord. Higher stretch rate must be applied
in the heat treating process for nylon 6 cords due to low modulus.
Nylon 6 loses tensile strength more than nylon 66 during high temper
ature cures. Accordingly, nylon 6 cord requires lower tire curing temper
atures or enough cooling when the tire is removed from the mold.
Nylon 6 and nylon 66 in common show flat spotting phenomena and
lower modulus than rayon and polyester. Many efforts have been made to
improve these properties, but all of them failed to achieve commercialized
materials.
The improvement of strength is one of the currently important prob
lems of nylon cord. The tenacity of recently developed nylon tire yarn has
reached ten grams per denier or higher [19].
Nylon monofilaments are currently under investigation for tire service
[19].
Production of nylon tire yarn:
To obtain high tensile strength, higher molecular weight polymer is
used and a higher draw ratio. Tensile strength increases progressively with
increasing molecular weight but dimensional stability of greige and
treated cord decrease at high temperature. Therefore, molecular weight
has to be selected to obtain the best balance of tensile strength and dimen
sional stability. Hot stretching at a higher temperature and other means to
improve dimensional cord stability [20-23] are adopted in production
processing of the yarn, e.g. multistep stretching, stretch and relax, etc.
Heat aging resistance of nylon cord:
Heat aging resistance is also an important property for tire service. Vari
ous additives are usually used in yarn production to improve aging resis
tance of the tire cord.
Thermal degradation of nylon cord in hot air decreases tensile strength.

CORD TREATING TEMP.CC } ]


NYLON 6 200 o
210 «
NYLON 66 210
230

130 150 170 19


TEMPERATURE CC )

FIGURE 2.6 Thermal shrinkage of nylon 6 and 66 tire cords as a function of temperature.
840 D/2 (47 x 47 turns/ 10cm.) (It).
TIRE CORD AND CORD TO RUBBER BONDING 45

This change becomes important at temperatures reached in heat setting


cords and in operating tires [24-26].
Decrease in tensile strength runs parallel to reduction in length of poly-
amide chains, reduction in the number of primary amine end groups, and
increase in the number of carboxyl end groups.
In order to improve heat and oxidation degradation properties, various
additives are used before or after polymerization. These additives are of
ten organic, e.g. some amines, some phenols, some haloaromatic acids,
diarylamine-ketone condensation products, 2-mercepto benzoimidazole,
and so on. Sometimes they are inorganic, e.g. phosphorous compounds,
and sometimes metals and their compounds, e.g. copper and its salts, or-
ganotin compounds, cobalt chelating compounds, and so on [24-25].
Heat treatment of nylon cord:
There are many reports concerned with heat treatment of nylon cords
and dipping and heat-stretching machines for tire cords or fabrics [ 1 8, 27-
32].
Nylon cords, or more generally, thermoplastic fiber cords, must be
dipped and heat stretched to improve dipped cord properties, e.g. to im
prove tensile properties, to reduce growth, (to increase Young's modulus
and to decrease creep), to improve dimensional stability, and to impose
cord-to-rubber bondability. A wide variety of single- and multiple-step
treatments have been adopted by various tire and fiber producers.
Cord or fabric treating conditions used by the tire industry at large vary
somewhat. Yet, within this variance, cords or fabrics produced under dif
ferent conditions all serve the precise purpose of the individual company.
Table 2.2 shows examples of variations that exist in conditions for treat
ing nylon 66 cords [30].
TABLE 2.2 Nylon 66 processing conditions for dying zone and single, dual or triple zone
treatment [30]
Tension, Residual
Company Time Temp. Ib/cord Stretch H20,% Tenacity, Ib
Drying conditions
A 1.5 min 270°F 1 1% 1% 24-26
B 2.5 min 240° F 0.75 1% 1% 24-26
C 3.5 min 180°F 0.5 1.5% 7-10% 23-25
Heat set conditions
A 20 sec 425°F 10 12-14% 0% 29-30
B 30 sec 4IO°F 10 10-12% 0% 28-29
C 45 sec 390°F 10 10-12% 0% 27-28
Normalizing conditions
A 20 sec 400°F ? 8% 0% 30
B 30 sec 370°F ? 8% 0% 30
C 30 sec 350° F ? 7% 0% 29
Reheat stretch conditions
A 20 sec 425°F ? 12-16% 0% 31.5
B Msec 410°F ? 12-15% 0% 31.0
Does not use triple zone treatment
Basis—840/2 nylon 66-Type 700.
Temperature levels for nylon 6 materials are approximately 30° to 40° below the nylon 66
values shown.
46 MECHANICS OF PNEUMATIC TIRES

We discuss the effect of treating conditions on the physical properties of


nylon 66 according to Du Font's data [27].
Figure 2.7a [27] shows the change in breaking strength that occurs in a
single step process as net stretch is increased.
Cold growth, heat shrinkage, and heat shrinkage tension are also depen
dent upon net stretch as shown in figure 2.7b [27]. Heat shrinkage in
creases with increasing net stretch up to a point, after which little further
change is seen. On the other hand, shrinkage tension increases in a fairly
linear manner with increasing net stretch at least to 20 percent net stretch.
Figure 2.7c [27] shows the results for a two-step process chosen to give
the same net stretch as the optimum one-step, with considerably better
heat shrinkage. Heat shrinkage tension does not improve. There are pat
ents on similar processes [33].

'0~~a T~* I 10 12 U If II 20
NET STKTCH (V.)

FIGURE 2.7 (a) Correlation between net stretch and breaking strength for nylon 66 cords
(single step stretching, (b) Correlation between net stretch and cord stability for nylon 66
(single step stretching), (c) Correlation between net stretch and cord stability for nylon 66
(two step stretching) [27],
TIRE CORD AND CORD TO RUBBER BONDING 47

TABLE 2.3. Correlation between stretch and thermal shrinkage for nylon 6 cord;
two step stretching [18, 32]
1st Stretch, 2nd Stretch, Total stretch, Elongation at Thermal
% % % 4.5 kg, % shrinkage, %
4 4 8 9.5 4.0
6 2 8 9.4 3.8
8 0 8 9.4 3.7
10 -2 8 9.3 3J
12 -4 8. 9.5 2.8
14 -6 8 9.6 2.7

Our similar results with nylon 6 are shown in figure 2.8 and table 2.3
[18, 32].
Good dimensional stability can be obtained by heat treating at high
temperature.
Drying conditions are also important. Loss of tensile strength in the
heat stretching process is marked when water in the cord has not been re
moved sufficiently in a dryer. This loss, however, can be avoided by adopt
ing a predip process (water dip prior to RFL dip [34]).
Among heat treating machines other than the usual types mentioned
above, a fluid bed process is worthy of mention [35]. This process, in
which small glass beads are used as the heating medium, was developed in
England.
Strength loss in tire curing:
In high temperature curing of nylon tires, particularly nylon 6 tires by
the Bag-O-Matic process, it is considered important to improve the ther
mal stability of the tire cord to withstand the curing conditions. Loss of
strength is more serious in the portion around the bead than at the crown.

e 12 190 zoo ao
STRETCH na TREATING TEMPERATURE CO

FIGURE 2.8 (a) Correlation between stretch and cord strength for nylon 6. (b) Correlation
between stretch and elongation at 4.5 kg. for nylon 6. (c) Correlation between stretch and
thermal shrinkage for nylon 6 [18-32]. (d) Correlation between treating temperature and
thermal shrinkage for nylon 6.
MOD/2, 47 X 47 turns/ 10 cm.
48 MECHANICS OF PNEUMATIC TIRES
NYLON 6 NYLON 66

113
5

13
2

&
CD
[1111111111 11111111
140 150 160 I'/O 180 140 150 160 170 180
CURING TEMPERATURE ("C )

MOLD OUT Wl THOUT COOLING


HOLDOUT AFTER COOLED DOWN TO 140° C

FIGURE 2.9 Strength loss of nylons in vulcanization.


MO D/2, 50 X 50 lurai/IO cm. 136).

Figure 2.9 indicates Pieper's result [36], in which strength loss of nylon 6
is compared with that of nylon 66. Strength loss of nylon 6 is greater than
that of nylon 66, particularly at high curing temperatures. However, adop
tion of cooling before removing from the mold is effective in avoiding the
strength loss. We also have examined loss of strength in tire cures with ny
lon 6, using a steel mold like Reegan and Sabos [37], to simulate the typi
cal actual shrinking behavior of cord which is expected in a tire curing
process [38]. The test diagrams of shrinkage are shown in figure 2.10.
Strength loss and thermal shrinkage in the curing process are illustrated in
figures 2.11 and 2.12, respectively.
From these results, it was concluded that strength loss of the cord is
largely to be attributed to rapid shrinkage when the work is removed from
the mold at high temperature and that a slow shrinkage of about 10 per-

METHOD A

D
E

PRESS OUTf
IkPOST STRETCH

FlOURE 2.10 Laboratory test diagrams of shrinkage used for evaluating strength loss in
curing.
TIRE CORD AND CORD TO RUBBER BONDING 49

IM 170 110 190 200


CURING TEMPERATURE ("c )
FIGURE 2.1 1 Strength retention of nylon 6 (T-78JS) after curing as a function of curing
temperature.
cent, which the cord suffers during the curing process, has little effect on
strength. Therefore, it is advisable to cool the cord and then remove it
from the mold. These phenomena have also been confirmed by actual tire
curing experiments. Figure 2.13 [39, 40] shows our experimental results in
which the strength loss of nylon 6 cords in tire curing is primarily affected
by the blowing-down temperature and cooling conditions, and is not re
lated to the maximum temperature used during the tire process.
Molecular mechanisms of strength loss with (rapid) thermal shrinkage
have been studied by various investigators. Dismore and Station [41] con
cluded from their study on nylon 66 that strength loss with thermal
shrinkage at high temperature is related to introduction of folded chains.

£,.

< 5

IM 170 180

CURING TEMP (°C)

FIGURE 2.12 Total shrinkage of nylon 6 (T-781S) through curing cycle.


50 MECHANICS OF PNEUMATIC TIRES

165 175 150 160 170


CURING TEMPERATURE ,°C BUOW DOWN TEMPERATURE ,°C
FIGURE 2.13 Relation between cord residual strength and tire curing conditions [40].

Fuji moto stated that there were good correlations between strength loss
with shrinkage at high temperature and the 002 lattice spacing of crystal
line regions or their change with thermal shrinkage [42]. Furthermore he
stated that when strength loss was great, a large number of filaments were
ruptured with sharp edges inclined approximately 35 degrees to the fiber
axis. There was a good correlation between strength loss and number of
bias breaks [42]. From these experimental results, he suggested that mech
anisms of strength loss, of thermal shrinkage and fatigue failure have
something in common, as will he mentioned later.
Todoki and Kawaguchi studied the melting behavior of nylon 6 yarns
by differential scanning calorimetry (DSC). They reported that y-ray irra
diation of nylon 6 yarns in gaseous acetylene provides an effective tech
nique for suppressing reorganization of nylon crystallites during the heat
ing process by introducing crosslinks into the amorphous region. The peak
temperature of the DSC melting curve thus obtained is considered to give
melting points strictly inherent to the original crystallites in the nylon 6
yarns [43].
We have studied the relationship between strength loss of nylon 6 cords
in tire curing and melting behavior of nylon 6 cords obtained by the
above-mentioned calorimetric method. The strength loss measured by the
method C shown in figure 2.10 is related to the peak temperature of the
DSC melting curve obtained by the Todoki- Kawaguchi method, as shown
in figure 2.14 [40]. Our results are also shown in figure 2.15, in which the
peak temperatures of heat-treated cords are shown as functions of the
peak temperatures of untreated yarns and conditions of cord heat-treat
ment [40].
When strength retentions are lower than 90 percent, the phenomena of
strength loss in tire curing are satisfactorily explained by these consid
erations. However, in the cases in which strength retentions are higher
than 90 percent, the phenomena become more complex.
TIRE CORD AND CORD TO RUBBER BONDING 51
100

Cord Treating Condition


200°C, 8% Stretch

195 200 205


MELTING PEAK TEMPERATURE ,°C
FIGURE 2.14 Relation between strength retention in model curing and melting peak
temperature of heat-treated nylon 6 cord (T-781S). Model curing method is the
method C in figure 2.10. Melting peak temperatures of DSC are obtained by the To-
doki-Kawaguchi method [40].

Mechanism offlatspotting:
It is generally assumed that the tire cord is primarily responsible for
flatspotting, in spite of other variables [44 45] in tire construction and
manufacture which affect flatspotting. Therefore, almost all articles which
deal with test methods or mechanisms of flatspotting are based on the vis-
coelastic behavior of the component fiber.
Some papers have dealt with laboratory tests on cords (or yarns) to sim
ulate the phenomenon of flatspotting in tires as affected by choice of tire
cord [46-54]. Other papers have been concerned with quantitative mea
surement of flatspotting in tires on indoor wheels [55-56].
Here, we will consider how tire cords behave in tire service. Figure 2.16
shows the deformation of a tire under load. Cord strain in the footprint is
smallerjhan in, other parts. All volume elements of the tire, however,
sJMUi an" equal time in the footprint during a revolution; their average
strain per revolution is the same. This situation no longer exists after the
tire stops. When the tire comes to rest under load after long running which
raised the tire temperature, the volume elements in the footprint cool to
ambient temperature under less strain than other elements. When the tire
starts to rotate again, persistence of this strain difference causes the flat-
spot. As the tire rolls and consequently is heated, this difference decreases
with time and eventually becomes negligible The qualitative phenomena
52 MECHANICS OF PNEUMATIC TIRES

Ul

180 190 200


PEAK TEMPERATURE OF UNTREATED YARN
FIGURE 2. 15 Relations between melting peak temperatures ofheat-treated nylon 6 cords and
untreated nylon 6 yams. Cord treating conditions are indicated [40].

can be investigated in the laboratory; the conditions which we examined


are illustrated in figure 2.17 [57].
Two cords are used in the experiments. First, they are given thermal
and mechanical conditioning steps to approximate vulcanization. There
after, these cords are given a thermal and mechanical history to simulate
the behavior of cords in actual tire operation; one cord represents a cord

FOOTPRINT

'iiiiuniiiiuifiiiiwiiiiiiii
FIGURE 2.16 Deformation of tire under load.
TIRE CORD AND CORD TO RUBBER BONDING 53
6 etO 1260 1260 1260 640 840 840 840 A "0
^
^^ 840
^
^" T~
,420
^_
C 105 25 105 25 IDS 25 105 105 25 25
MIN 60 10 60 30 I6HR 10 20 5 20 10 SEC
RUNNI N&—• •PAR KING - •-RUNNING

FIGURE 2.17 Laboratory testing methods for flatspot index of cords [57].
A. I in Pont method.
B. Modification of Goodyear method.
C. Modification of Allied method.

located in the footprint when the tire was brought to rest and the second
cord represents a cord in other parts of the tire.
The difference between A and B (see figure 2.17) is an index of the mag
nitude of flatspotting.
Finally, to simulate the run-out of flatspotting, both cords are reheated.
Tippetts explained the magnitude of flatspotting in terms of the differ
ence in Young's modulus of the cord at tire operating temperature and at
room temperature [47]. He concluded that magnitude of flatspotting is
proportional to (1/MH) X [1 - (MH/MR)], where MH is Young's modulus
at tire operating temperature and MR is Young's modulus at room temper
ature.
Flatspotting becomes large when the operating temperature exceeds the
glass transition temperature Tg of the cord, because then (1/MH) [1 -
(MH/MR)] becomes large rapidly.
Relationships between flatspot index and tire operating temperature,
which were examined by Papero and co-workers, are illustrated in figure
2.18 [49]. In both cases, nylon and polyester flatspot indices increase rap
idly when tire operating temperature exceeds the Tg of the cord.
Flatspot index decreases as cord Tg increases, as shown in figure 2.19.
[57]. Moisture regain of the cord also has important effects on the flatspot
index of water sensitive cord because absorbed moisture lowers the Tg of
the cord, figure 2.20 [57]. The flatspot index rapidly increases, then passes
through a maximum and finally decreases with moisture content. Reduced
flatspot nylon, modified by blending of aromatic polyamide, is claimed to
exhibit good flatspotting resistance at low humidity but loses its merit at
high humidity.

*o 60 ao no no wo
TEMPERATURE ("C )

FIGURE 2.18 Flatspot index as a Junction of running temperature [49].


54 MECHANICS OF PNEUMATIC TIRES

? NYLON 60R66-AROMAT1C NYLON


OS COMBINATION
OS - • NYLON 6- PET COMBINATION

cu
100 110 120 130 140 150
' tonSmoK ( C )

FIGURE 2.19 Flatspot index as a function ofT (tan oml^ (as measure ofTgJ [57].

1.0 1.3 2.7


MOISTURE REGAIN (V.)

FIGURE 2.20 Flatspot index as a function of moisture regain [57].

TABLE 2.4. Reduced flatspot nylons


Name Method Literature
Crosslinked nylon Vapor phase treatment with diisocyanate [58]
Cyanuric chloride treatment on nylon 6 [59]
N-44 (DuPont) Melt blend yarn of nylon 66 or 6 and [51,60-62]
aromatic polyamide
X-88 (Monsanto) Nylon 66 modified by aromatic [63]
polyamide based on terephthalic acid
EF-121 (Allied) Melt blend yarn of nylon 6 and 30% [49, 52, 60]
polyethylenterephthalate
NF-20 (Firestone) Block copolymer of polyamide and [60,64]
polyester
Merged fiber Merged fiber of nylon-polyester [65]
Nomex [66]
6T fiber M
TIRE CORD AND CORD TO RUBBER BONDING 55

Many efforts have been made to reduce flatspotting of nylon cords as


shown in Table 2.4, but all of them have failed to achieve commercialized
materials.

2.1.4. Polyester Cord [68-72, 78]


Almost all polyester yarns used in the tire industry are filament yarns of
polyethylene terephthalate.
As rayon and polyamide dominate the tire cord market an improved re
inforcing material is expected to meet the requirements imposed by in
creasing severity of tire operating conditions. A fiber, to be a satisfactory
tire cord material, must possess a rather special, balanced combination of
properties.
Polyester's high modulus and low elongation reduce tire deformation
and growth under service conditions and lead to better high speed per
formance and tread wear, reduced tread cracking and better steering char
acteristics. In addition, polyester's dimensional stability allows manufac
turing a more uniform tire.
High strength or tenacity is required to provide adequate carcass
strength and in this respect, nylon is superior to current commercial poly
ester, which in turn is superior to rayon.
Polyester is considerably superior to rayon in fatigue resistance. In sum
mary, polyester yarns have a better balance of fundamental properties for
efficient use in modern tires than does rayon, and replaced a large portion
of rayon cord.
Manufacture ofpolyester tire yarn:
Differences in production processes for high tenacity polyester yarn
from those for regular polyester yarns are explained below. To improve
the tenacity of cords, molecular weight of yarns and draw ratio in the pro
duction of the yarn are important factors. Intrinsic viscosity, IV, (a mea
sure of molecular weight) of current tire yarn is 0.8 to 0.95, while that of
regular polyester yarn is under 0.6. Draw ratio in tire yarn production is
also much higher than that in regular yarn production.
Draw ratio in the production of the yarn has a large effect on both yarn
tenacity and the yarn-to-cord tenacity conversion efficiency [68]. Higher
draw ratio results in higher tenacity of yarn and lower conversion effi
ciency. This decrease in conversion efficiency eventually reaches such pro
portions (as yarn tenacity is increased by increasing yarn draw ratio) that
cord tenacity actually begins to decrease in spite of higher starting yarn
tenacities; see figure 2.21 [40].
The slope of the upper part (near the breaking point) of the stress-strain
curve significantly affects yarn-to-cord conversion efficiency [40, 68, 73].
The higher the slope of the upper part of the stress-strain curve the lower
becomes the conversion efficiency. Figure 2.22 shows our experimental re
sults [34]. This tendency is explained in terms of homogeneity of stress dis
tribution in the cord by twist theory [74]. The slope of the upper part of
the stress-strain curve is greatly affected by the draw ratio in the produc
tion of yarn, but other factors are also significant.
Tenacity of the cords can be increased without increasing the slope of
the final part of the stress-strain curve or decreasing the yarn-to-cord con
version efficiency by increasing the molecular weight.
Using a suitable lubricant, which decreases the static friction coefficient
MECHANICS OF PNEUMATIC TIRES
i r
_ 10

8
8
I
Q
3
o 8

>7
§
£ 6

5.5 60 6.5 7.0


MECHANICAL DRAW RATIO
A: IOOOD/2 40Zx40S(turn/10cm)
B: " 45Zx45S
C: 5IZ x51s

FIGURE 2.21 Relation between draw ratio of yarn and tenacity of yam and raw cord with
various twists [40].

100
CORD CONSTRUCTION 1 100 D/2

90
40x40

80

70

I 2 3
ATj.5 g/d , %

FIGURE 2.22 Relation between &T25g/d and yarn-to-cord tenacity conversion efficiency.
TIRE CORD AND CORD TO RUBBER BONDING 57

of the yarn, is also effective for improving the yam-to-cord conversion ef


ficiency, as in the case of rayon.
Heat treatment ofpolyester cord:
Heat treating conditions of polyester cords have a pronounced effect on
their properties. The heat treatment temperature of polyester cord is con
siderably higher than that of nylon or rayon. Stretch rate for polyester is
considerably lower than that for nylon. Typical heat treating conditions of
polyester cords are compared with those of nylon and rayon cords in table
2.5 [38]. In the case of polyester cord, fatigue resistance is particularly sen
sitive to the heat treatment.
Aitken and co-workers reported that although stretching ratio affects
the fatigue rating, treating temperature is the single most important factor
in achieving high fatigue rating with polyester tire cord, table 2.6 [68].
They concluded from their results (shown in table 2.7 [68]) that strength
retention of the cords after heat treatment is best for the highest IV sam
ples, which increases the advantage of high IV material over low IV mate
rial. This effect is most marked at high heat-treating temperatures. The
higher the cord IV, the higher the fatigue rating which can be achieved. A
reduction in Mallory rating occurs at all IV levels when treating temper
ature are taken to 490° F. High IV cords suffer less than low IV cords in
this report, however.
At 450° F, the Mallory ratings of the low IV samples exceed those of the
high IV samples. At the two higher temperatures investigated, 475 °F and
490° F, this effect is reversed.
In our experiments dealing with higher molecular weight (0.85 and 0.95
IV) yarn, rather high temperature heat treating may give a good fatigue
rating and good dimensional stability but results in low strength cord and
in a stiff fabric, which may cause difficulties during tire building [34].
The strength loss of polyester cord in high temperature heat treatment,
which is usually about 10 percent of initial strength, is similar to the
strength loss of nylon cord in high temperature tire curing.
In the case of heat treatment of polyester cords, the remarkable change
of crystal line structure is observed by small angle x-ray scattering, but no
change in IV is observed. [40].
To improve strength retention in high temperature heat treatment, the
diethylene glycol content must be reduced to a low level. Diethylene gly-
col content affects the softening temperature of polyester yarn [75].
In the case of the polyester tire, the strength loss of cord in tire curing is
not serious, and increases in strength are sometimes observed.

TABLE 2.5. Comparison of Typical Heat Treating Conditionsfor Rayon, Nylon, and
Polyester (38]
Set temperature Net stretch
op
°C %
Passenger tire
Rayon 155-165 310-329 1.5-3.0
Nylon 6 205-210 401-410 5.0-9.0
Nylon 66 220-230 428-446 4.0-7.0
Polyester 235-245 455-473 0-4.0
Truck tire
Nylon 6 207-212 404.6-413.6 6.0-10.0
MECHANICS OF PNEUMATIC TIRES

TABLE 2.6. Optimization of heat-treating conditions for polyester cord [68]


Heat treatment
Mallory
temperature,* First oven Second oven fatigue
oF stretch, % stretch, % rating, ke
A 400 3 0 11
B 400 5 -2 12
C 450 3 +2 24
D 450 5 -2 48
E 475 4 0 134
F 475 4 -3 196
* 2/1 100, 13 x 13 cord made from Type 24 "Terylene" yam. Dwell time 90 sec in all cases.

Durability ofpolyester tires [78]:


Polyester tire cord is already used broadly in the field of passenger car
tires and is about to come into use in the field of truck tires. Compared to
passenger car tires, truck tires generate more heat under service conditions
and are subjected to more severe operating conditions. It is clear that the
passenger car tire will also meet more severe operating conditions with the
development of expressways and high-speed cars.
Table 2.8a shows the results of our laboratory indoor test using the poly
ester bias tire of 7.00-14 4PR (actual 4 plies). Tires A and B failed due to
cord break-up after running for the distance of 7,293 km and 4,942 km, re
spectively. The cords taken out of these tires retained 85 to 94 percent of
their initial strength at the crown, showing a favorable strength retention
rate. On the other hand, they retained only 20 to 64 percent of their initial
strength at the shoulder, which is placed under the most severe conditions.
On measuring the intrinsic viscosity TJ of the respective cords, it was
found that the viscosity of the cord at the crown remained almost un
changed from the condition prior to running, while that at the shoulder
was down to 60 percent of its original value.
Measurements were made on the air temperature contained in the tires
using a slip ring and an AC thermocuple. The contained air temperature
reached the figures shown in Table 2.8b in about one hour after the start
of running. From the fact that the temperature reached 130°C, it is easily
imagined that the temperature would be quite high at the shoulder.
These decreases in strength and ?/ values are considered due not only
to the physical deterioration under extension and compression but also

TABLE 2.7. Effect ofIV and treating conditions on treated cord propertiesfor polyester [68]
Greige Treating Breaking Mallory
breaking temperature, load fatigue
load, Ib op retained, % rating, ke
Yarn IV
490 76 8
0.60 28 475 86 132
450 92 122
490 81 21
0.69 30 475 87 159
450 95 61
490 85 97
0.73 31 475 91 175
450 95 23
TIRE CORD AND CORD TO RUBBER BONDING 59

TABLE 2.8a. Results of indoor wheel lest on polyester tires [83].


Cord taken out oftires
Running Intrinsic
life Strength Viscosity
(km) Breakage (kg) % ft)
Before Tire A 15.5 (100) 0.76
Running Tire B 15.2 (100) 0.76
crown Cord Break- 13.2-14.5 (85-94) 0.79
Tire A 7791 up
After shoulder 6.0-10.0 (39-64) 0.52
Running crown 13.2-14.3 (87-94) 0.75
Tire R 4947 CnrA Brealr-
shoulder up 3.0-8.0 (20-53) 0.43
Test conditions: inflation pressure: 1.4 kg/cm2; tire load: 0-1400 km (warm-up running) 425
kg, 1400 km-failure (test running) 625 kg; speed: 60 km/h; ambient temperature: 38 ± 4°C;
test wheel: 170 cm in diameter.

TABLE 2.8b Contained air temperature [83].


0-1400 km 1400 km-failure
(Warm-up Running) (Test Running)
Tire A 94- 97 °C 123-125°C
TireB 96-102°C 126-130°C

largely to the chemical deterioration by hydrolysis and aminolysis caused


by heat generation.
These problems are closely related to the terminal COOH group quan
tity of polyester [75-77]. The heat aging resistance test by the glass tube
method was conducted in our laboratory on samples different in the
COOH end group content. The results are shown in figure 2.23. It is obvi-
100

90

Dipped Cord
80

70

60-

50 -

10 15 20 25 30 35
COOH END GROUP CONTENT

FIGURE 2.23 Relation between the COOH end group content and heat resistance (glass tube
method) [83].
60 MECHANICS OF PNEUMATIC TIRES

ous that heat resistance can be improved by decreasing the terminal


COOH group quantity.
From the facts mentioned above, we are ready to develop a process
which decreases the COOH end group content by various means in order
to manufacture advanced heat resistance yarns. A polyester tire cord with
the COOH end group content of less than 5eq/1o6g is now experimentally
available in our laboratory.
Several factors in rubber are also important in the chemical degradation
of polyester cord and the adhesion of polyester cord to rubber.
Carcass stocks which minimize or eliminate the presence of amines re
tain good performance in polyester tires in severe tests. These stocks con
tain vulcanization accelerators that do not liberate amines or additives to
inhibit the deleterious effect of amines present as impuritites. [78-83].

2.1.5. Steel Wire Cord [84-90|


Steel wires used in the tire industry are manufactured by the cold draw
ing process. Typical specification of high carbon steel for tire cord is [4]:
Carbon 0.7%
Manganese 0.5%
Silicon 0.3%
Chromium 0.05%
Copper 0.02%
Sulfur 0.03%
Phosphorus 0.03%
Steel rods are drawn in two steps through dies of diamond or tungsten car-x
bide to reduce the diameter to about 1/8 of its initial diameter. Between
and after the drawing steps the wire is heat-treated to change the grain
structure (called patenting).
The wire is then brass plated. The brass plating enhances the adhesion
ot wire cord to rubber and facilitates drawinejQ_fine diameter. The wire is
drawn finally to a diameter in the range "ofM^O
The drawn wire filaments are then combinedTo form a strand. Several
strands are combined to obtain the final wire tire cord. Typical wire tire
cord constructions are illustrated in figure 2.24.
The wire cords are wound on convenient spools and usually subjected
to a rubber topping process (calendering) without passing through the
weaving and the adhesive treating process [4]. Steel cords are mainly used
in truck, bus, and off-the-road tires. In the case of passenger car tires, steel
cords are mainly used for the belt cords of radial tires.
Monsanto has attempted to develop the melt spinning process of steel
filaments. It is reported that melt-spun steel filaments have fine diameters,
so fatigue resistance and processability are improved [91].
According to patent specifications of Monsanto, the outline of the proc
ess is [92]:
Steel alloy (consisting of 90 percent carbon steel and 10 percent alumi
num, for example) is melted in a crucible in vacuum and extruded under
argon gas pressure through a fine diameter orifice into a chamber with at
mosphere containing ammonium or oxygen and nitrogen. By exposure of
the extruded melted alloy to such an atmosphere, a stabilizing film is gen
erated on its surface. Filaments of about 100 microns diameter are ob-
(-)„ GOOl"
TIRE CORD AND CORD TO RUBBER BONDING 61

(7X3) (5X7H1X3)
FOR PASSENGER CAR FOR TRUCK

FIGURE 2.24 Typical cord constructions of steel wires.

tained. These must be heated to homogenize their structure to give supe


rior mechanical properties.
2.1.6. Fiber Glass Cords [93-104, 114-116]
Simply stated, the most important problem of fiber glass cord is effec
tive use in mu lt i filament structures of superior single filament properties
shown in table 2.9. [93]. Characteristically, single-filament properties are
reducedby more than one-half when formed into strands or yarns. Early
developments showed that excellent flex life of single-filaments dis
appeared when they were combined as a cord.
For years, many attempts have been made to solve these problems. Phe
nol formaldehyde, elastomeric materials, and more recently, some of the
organo-silicon compound coatings were found to be effective to improve
these properties.
Some inorganic coatings such as metals were also effective in improving
abrasion resistance and flex life of multifilament glass fiber structures.
However, lack of rubber-to-cord adhesion and insufficient filament-to-fil
ament stress transfer resulted in unsatisfactory performance.
Various surface finishes were investigated to improve glass-to-rubber
adhesion. A combination system of a compatible surface finish and an
RFL was also developed as an impregnation material to improve filament
to filament abrasion. With these developments it became possible to main
tain single filament properties in impregnated multifilament glass cord; see
table 2.10 [98].

TABLE 2.9 Physical properties of single filaments of glass and of organic fibers [93]
Glass Rayon Nylon Polyester
ECC' T-130 T-A05 T-52
Tensile strength ultimate, psi 500,000 79,000 126,00 139,000
Tenacity at break gpd 1532 405 865 789
Ultimate elongation, % 4.76 11.74 19.93 14.92
Modulus, gpd 322 35 43 53
1000 psi .. . 10,500 680 630 940
Toughness, gpd 0.365 0.312 1.05 0.720
psi .... 11,900 6,100 15,300 12,800
• Sp. gr., 2.55; elastic recovery, 100%; softening point, 1555°F; coefficient of thermal ex
pansion (°F), 2.8 x 10.6; and water absorbency, 0.3%.
62 MECHANICS OF PNEUMATIC TIRES

TABLE 2.10. Physical properties oj'cords of glass and organicfibers [98]


Glass Glass Rayon Nylon Polyester
ECG-150 ECG-150 T-130 T-130 T-52
10-1A/0 10-1A/3 1650/2 840/2 1100/2
Tensile strength
(ultimate), psi 407,000 365,000 94,000 122,000 104,000
Tenacity at break,
gpd 12.47 11.17 4.79 8.37 5.86
Ultimate elongation,
% 4 83 4.84 9.8 19.3 18.5
Toughness gpd 0302 0.254 0.296 0.696 0.560
osi 9900 8,300 5,800 10,200 9,900
Modulus, gpd 259 231 49 43 32
1000 psi 8,450 7,540 960 630 570
Breaking strength, Ib.. 79 219 39.1 33.2 32.1
Impact resistance,
(ft/lb x 10~4)/denier 395 2.75 1.87 4.08 3.41
Diameter, mils 17 35 26 21 24
Number of filaments 2,040 6,120 8,000 280 500
Specific gravity (fiber) 2.55 2.55 1.53 1.14 1.38

2. 1 1 shows physical properties of the improved glass tire cord


'^<ttycor", with high RFL pick up. Filaments are bound as a bundle with
lant (RFL type adhesive), and at the same time separated from
each other so as to avoid inter-fiber abrasion. Such an improved coating
process results in higher strength and resistance to flex fatigue.
Continuous filament fiber glass is manufactured by the melt spinning
process. A typical composition of glass for tire cord is:
Silicone dioxide 53%
Calcium oxide 21%
Aluminum oxide 15%
Boron oxide 9%
Magnesium oxide 0.3%
Other oxide 1.7%
This is called "E"-glass, a lime-alumina-borosilicate glass consisting of
a three dimensional silica network containing oxides. Raw glass materials
are melted and extruded through orifices into filaments. They are solidi-

TABLE 2.11. Physical properties ofimproved glassfiber cord 'Hycor" [103]


G-75 5/0 1.5S Hycor 18% DPU
Yardage yd/lb 1150 1260
Dip Pick- up,' , 30 18
Cord Diameter mils 23 20
Tex. gm/ 1000 M 432 393
Breaking Strength, Ihs 66 60
Glass Tenacity, gm/den 9.8 8.9
In-Rubber. Strength, Ib Adhesion, Ib 74.9 72.5
U-Pull (Goodyear) 27.0 23.0
H-Pull (BFG) 16.0 13.0
T-Pull (Firestone) 15.0 13.0
Strip (Uniroyal) 36.0 33.0
Strip, Room Temp. 57.0 45.5
250°F 40.6 35.6
Scott Flex. K.C 600 250
TIRE CORD AND CORD TO RUBBER BONDING 63

tied by a water quench, and then impregnants are applied as mentioned


above [94].
Fiber glass cord is mainly used for the belt cord of belted-bias tires.
Recently, a fiber glass radial tire has been developed consisting of a
one-ply fiber glass carcass and two plies of fiber glass belt.
2.1.7. Wholly Aromatic Polyamides [105]
The characteristics of the wholly aromatic polyamide fibers are ex
tremely high modulus, high strength, and mechanical and chemical stabil
ity at high temperature. These properties are derived from the stiffness of
its specific molecular structures.
Du Pont's "KEVLAR" (called Fiber B initially). Monsanto's "X-500",
AKZO"S ENKATHERM and TERLON (Soviet) are known as wholly
aromatic polyamide fibers. "KEVLAR" is the fiber from poly-para-
phenylene terephthalamide. Its melting point is over 570°C, with a glass
transition point over 300° C (higher than the melting point of PET or ny
lon 66). "KEVLAR" has a modulus of about 300 grams per denier and a
tensile strength of about 15 grams per denier at 200°C. X-500, a fiber from
polyamidehydrazide, also has a high melting point.
These fibers are made by the dry and wet spinning process. Du Pont has
proposed to use a generic name "ARAMID" for "KEVLAR" and similar
aromatic polyamides. Representative data for the fiber properties of
"KEVLAR" are shown in figures 2.25 and 2.26 and table 2.12.
"KEVLAR" was initially tested as a belt cord for radial tires. Wilfong
and co-workers reported that "KEVLAR" belted radial tires showed
equivalent durability to wire belted controls at a 1 to 5 replacement ratio.
The results of wheel tests are shown in tables 2.13 and 2.14.
2.1.8. Miscellaneous cords [106-108, 114, 116]
In addition to the cords mentioned above; PVA, nylon monofilament,
and macrofilament polyester have been investigated as tire reinforcing
materials.
I i 1

24 -
Fiber B
§20 ARAM ID

1.6 _
Nylon
« 12 - / Polyester
a B n D GI°»
S 8
u 1 Rayon
P
4 0 Win>

II 1 1 1 1
100 200 300 400 500 600 700
MODULUS (g/d)

FIGURE 2.25 Tire fiber tensile strength-modulus relationships [105].


64 MECHANICS OF PNEUMATIC TIRES

0 50 100 150 200 0 50 100 150 200


TEMPERATURE, °C

FIGURE 2.26 The effect of temperature on tensile and modulus properties [105].

TABLE 2.12. Dimensional stability of'dipped tire cords [105]


160°C
Fiber Shrinkage (%) Growth (%)• Creep (%)** DryTg(0C)
T-728 Nylon 6.8 4.8 0.4 50
T-68 Dacron 6.0 XI 0.3 75
Rayon 0 4.9 1.4
Fiber B 0-0.2 0.5 <0.03 300-1-
Wire 0 0.7 <0.03
Glass 0 0.5 <0.03
* Test conditions: 1 g/d. 30 min 24°C
•• Growth between 30 sec and 30 min

TABLE 2. 13. Fatigue resistance ofcarcass cord in 4-ply 7. 75-14 bias tires [105]
Twist Tenacity Elongation Tire Fatigue*
Fiber Multiplier (g/d) (%) Strength loss (%)
T-728 Nylon 6.6 8.7 21 1-10
T-Dacron 6.6 7.2 15 10-20
Fiber B 8 15 8-16
* Wheel test: 3000 miles at 21% dynamic deflection: 18 psi inflation pressure at tire run
ning temperature: 85 mph.

TABLE 2. 14. Durability ofradial tires (polyester carcass) [105]


Belt Fiber
Steel Wire Fiber B
Belt Fiber, Ib 2.5 0.5
Relative Tread life (%)* 350 350
Wheel Test Failures
Step- Load Endurance (%)
(side wall failures at ply turnups) 130 130
High speed (miles/hr) 100 100-105
(105-110)*
Cleated wheel (miles)f 900 1500
(3500)**
* Relative to 4-ply bias rayon controls. ARA Severe XHA-1 test route, 28,000 miles.
** Edges of Fiber B belts were folded to remove cut edges from high stress shoulder zone,
a difficult operation for wire because of its stiffness.
† 12 rounded cleats, 3/4 inch high by 2j inch wide, bolted of various orientations to sur
face of the 5.6-ft-diameter test wheel. Tire inflated to 24 psig and run 40 miles per hour.
TIRE CORD AND CORD TO RUBBER BONDING 65

Vinylon (Kuraray's PVA fiber) is used to some extent in bicycle tires


but not in automobile tires because of inferior heat aging resistance. Ku-
raray has developed a new PVA filament yam of high modulus and high
tenacity, which is manufactured by wet spinning process and has good
thermal stability [107, 108].
Hannell reported that the use of nylon in monofilament form reduced
the cost of tire cord processing prior to tire building [19].
Dunlop U.K. announced that they were introducing an all polyester
radial tire utilizing polyester macrofilament yarn in the belt. [106].

2.1.9. Cordless tires [109, 110]


Experiments on the cordless tire by Firestone showed that the cast tire
can pass the DOT wheel tests with strength in reserve. In addition, in
static and high speed running tests, cast tires exhibit desirable qualities for
practical use. Cordless tires, if realized, will eliminate a large portion of
conventional tire facilities such as banbury mills, tire building machines,
calenders and so on. The most important problem in the cordless tire is the
specification of the elastomer. Firestone has not published details of the
materials used in their experimental tires.
2.1.10. Comparative Analysis of Various Tire Cords [11-116, 40]
A number of articles have appeared which are concerned with com
parative analysis of various tire cords. The results which are summarized
in table 2.15 were reported by Schroeder and Prettyman [111]. In increas
ing order of tenacity and breaking elongation are rayon, polyester, modi
fied nylon and nylon. Rayon and polyester have higher initial moduli than
do either of the nylons. Rayon possesses greater thermal stability than pol-

TABLE 2.15. Physical properties of various tire cords [11 1]


Rayon Nylon
(current) Nylon (modified) Polyester
1650 D/2, 840 D/2, 840 D/2,
12 X 12 12 x 12 12 x 12
2200 D/3, 1260 D/2, 840 D/2, 1260 D/2,
Cord construction 9.5 X 8.5 10 x 10 12X12 10 X 10
Tenacity gpd 45 7 6.5 6
9 21 15 13
60 27 40 70
Creep, % 3 4 3 2
Shrinkage, % 06 6 7 7
Flatspot Index, mils 50 170 US 50
Impact toughness,
erg/em D 225 525 400 375
C-F fatigue, min 400 1400 600-1200 500
Adhesion Ibs/in 100 90 70 65
70 95 70-95 70
Esthetic rating of tires
Flatspot Par Even
Roughness Par 4- + ». Even
Ride Par Even
Par
Noise brick road Par
Handling Par Even Even Even
66 MECHANICS OF PNEUMATIC TIRES

TABLE 2.16. Characteristics ofsuper-goal tire cord 1111]


Level
Property suggested Tire-responses reduced
Tenacity: 75°F to 400/F High Impact breaks, heat breaks, heat buildup.
high speed failure.
Elong, at break High Impact breaks.
Tensile modulus High Flatspotting, growth, high speed failure.
to
medium Impact breaks, flex failure, roughness.
Bending modulus Low Flex failure, roughness.
Creep Low Growth, tread wear, sidewall checking,
tread cracking, high speed fail re.
flatspotting, tire dimension and
uniformity variations.
Thermal shrinkage Low Tread concavity, bead distortion
to uniformity variations.
high high speed failure.
Flatspotting index Low Flatspotting.
Impact toughness High Impact breaks.
Flex fatigue resistance High Flex breaks.
Adhesion High Tread and ply separation.
Heat aging resistance High Impact breaks heat breaks, flex breaks.

yester and nylon. Rayon and polyester have higher flatspotting resistance
than nylons. Nylon shows a definite superiority for both impact toughness
and compression-flex fatigue resistance. Adhesion tests place rayon on
top, followed in order by nylon, modified nylon, and polyester, although
all are at a high level with modern dip systems.
We can conclude from these results that any current commercial cord
cannot satisfy all cord properties required.
Schroeder and Prettyman also indicated the desired directions for cord
properties as shown in table 2.16.
Ebert [8] also compared various textile cords with steel cord as the rein
forcing material for the belt ply of radial tires, see table 2.17.
Extra high modulus rayon and steel are promising materials for this
purpose from the viewpoint of initial modulus.
Some polyesters also seem to be acceptable.
Wilfong also compared various cords with "KEVLAR" cord as pre
viously shown in figures 2.25 and 2.26 and tables 2.12 and 2.13.

TABLE 2.17. Physical properties of cords for radial tire construction [8]
1Elongation
at 20% of
breaking Adhesive Impact Bending Bending Heat
load power energy life stiffness growth
% tons cm kg cycles cmg %
Steel 1 0.4 7 230 40 1,400 0.2
EHM 2 0.7 16 450 1,900 80 0.3
3 0.5 IS 700 1,000 60 0.3
Rayon 4 0.5 18 730 1,500 100 0.4
5 0.9 14 1,700 10,000 25 0.9
Polyester 6 1.0 1 800 4200 700 05
7 1.1 20 1,600 5,200 140 0.5
8 1.9 14 1,400 5,600 30 0.5
Nylon 9 5.0 19 1,800 26,500 17 0.2
TIRE CORD AND CORD TO RUBBER BONDING 67

i- i_
•9
111 !J \ 1 IIJSS8
,v , II I 1 11
u
8 88
iS —K U-X 1 < 8K IIj
Q1

i
|
^
| 0 0
f
i0.8u .•a 7 i
— &
a
o
&
:§ i i 1
i
& I |Ecd iId
1

g.0 ^'i
E S Q f ^J \

|
a> e
t
H

fl I t1 f
1
Q.
Z |s K 1 % I
| Rayon,— —Steel~
ii 1 j »*g
5 — C iii
if' | —Steel,
g
KI./I™. 1
i

j!j
X >> I'll r^i ^ :z
M,.|™.
1 1^ ! i JI Ij

1 4
•o -
"2
•38 a |o
•«
1 \\ _i n> jj
g .a s s
'1bymadsoLowsohirsitnukraege jadso:sbymLowhoirsitn>lukraegeHighmeltingpoint.Lowmoistui
c c .5
| •o : 0
gd>- 1torspeenergyup?<§turebonding,Highcord-soto-ruber
?•>> •S
2.18tawteion Tire
bRtire
elprof Desirable
Cor adHeymino
& raodlaytiocn (high
High point
melting
gHigh
&translass snhirfionrkmaigtey
uLow
(good
I
High
gmroalass
t& dnsuiltuiosn dkur1owability
lHigh
&
strength
Properties
High
mlodcreep
& owulus
performance/cost
High
gLow
heat
eneration
dynamic
Low
loss
Softnes
cord
of High
modulus High
modulus
modulus
Low High
modulus
Toughnes resistance

a
H
5
1
!P ^

sg «
t %at
\N 8 •58
.§ fffr* £ \
Illfj 1 Cr \ te°pi
Highmpe—rature Is
roblems Selvage
loosei— in
cord, Proces a—bility in
tire
Unicifo—rmity Strength
H 8 ^ H ^
produ
Pin
tire tire ofcore— air
loss blow
blist—er,
Air Onsur— rough
road
smooth surfa
road
On— Seltorque
f-al—igning
£ Apersotphertices pCro pnertinegs Corpower
ne—ring consum—ption
Fuel
resista—nce
Bruise
£ Retread—ability
f= |1| ||||j spotting—
Flat Treadwear—
price—
Tire
Economics

\$ ^^
68 MECHANICS OF PNEUMATIC TIRES

TABLE 2. 19 Usage pattern of tire cords [40]


Bias Tire Radial Tire
Tire Carcass Cord Belt Cord
Tire Small* Large* Small' Large' Small* Large*
Cord Tire Tire Tire Tire Tire Tire
Rayon O O ©

Nylon © © © O
N66 N6
O
Polyester © © 0 O
Steel O © © ©
Glass Fiber © •••
Aromatic polyamide 0 0 0 O © O
©Widely used.
OSometimes used.
•Small Tire: for passenger car & light truck
Large Tire: for heavy truck & bus
" High modulus modified polyester
••• For bias-belted tire

We have summarized the desirable properties of tire cords from the


view point of the tire production process, durability, safety properties, aes
thetic properties, cornering properties, and economics in Table 2.18 [40].
The usage pattern of various tire cords is summarized in Table 2.19 [40].

2.1.11. Impact Resistance of Tire Cords [117-121)


Impact failure of tires and tire cords, which will be considered here, is
one of several failures that may occur in high speed operations. In general,
impact resistance of a tire is determined either by road or simulated road
tests, or has been inferred from laboratory tests at low speeds and room
temperature.
Road tests are difficult and expensive and are limited in their ability to
provide specific technical information for the design and modification of
the design of tires. It is expected from the well-known behavior of poly
mers that deflections, breaking forces, and breaking energies under high
speed, high temperature conditions will differ greatly from those deter
mined at low speeds and room temperature.
Laboratory impact tests of tires and tire cords are important as a design
base for modern high speed tires.
Many superior laboratory methods of testing impact resistance of tires
and tire cords have been proposed but we will not discuss them in detail
here.
Hall [117] investigated the impact behavior of various fibers at a very
high rate of straining, 330 sec."', and at a normal rate, 8.3 × 10-3 sec.-1, for
a range of lightly twisted yarns covering all commercially important fiber
forming polymers. Details of his experimental results are given in table
2.20. From these data, we see that at the high rate, breaking stress is al
ways greater and with one exception, breaking extension less than at nor
mal rates. Energy to rupture increases with rate for wet spun fibers and
TIRE CORD AND CORD TO RUBBER BONDING

II
I I I I ill I I •i s^
— — H — .-, r* —• ill
|2^
8 &—
ig'g
*) O ""
•9«l

ill |!"
5|a *:M
SiJU

111
I !, J
ff r §2-^
07 afi-5
X g
06 -^^
•Us
o~ ^888888888888888888 888
•a-

OQOOgOOQQQOOOQOOOOO
<
8I1
19

ll|
esf
*M
^8
^ |||
ft li lii
8 — »?KN — — r ^3
ri |3 |
a
>>fe X
z" IP
»

I |l
1 5-

70 MECHANICS OF PNEUMATIC TIRES

with one exception and decreases for thermoplastic fibers. Similar experi
ments performed with three of the yarns highly twisted, the data of which
are given in table 2.21 and 2.22 [117] ranking of the yarns according to a
particular property could be altered by insertion of twist.
Lothrop [118] determined the tensile properties of rayon and nylon
cords as a function of temperature and rate of extension. His testing
equipment was designed and built to be capable of measuring the proper
ties of tire cords over a range of temperature from 75° to 300°F at rates of
extension of 1000 percent to 6000 percent/sec. (10 to 60 sec.-1).
His complete data for treated rayon and nylon cords are summarized in
table 2.23. Characteristically, for both oven-dried rayon and nylon, break
ing strength decreases as temperature increases at a given rate of exten
sion. Likewise, for a given temperature, breaking strength increases as rate
of extension increases. For oven-dried nylon cord, breaking elongation is
reasonably constant throughout the complete range of testing conditions.
Oven-dried rayon, on the other hand, shows an increase in breaking elon
gation with increase in temperature, as well as with increase in testing
speed so that the maximum breaking elongation was observed at a tem
perature of 300°F and a rate of extension of 6000 percent/sec. (60 sec.-1).
Breaking energy of the oven-dried nylon cord decreases with increasing
rate of extension at the lower test temperature but not at the higher test
temperatures.
Breaking energy of oven-dried rayon decreases with increasing temper
ature at low rate of extension but increases with increasing temperature at
high rates of extension.
Recently, Lothrop published another report [119] which confirmed the
above results by tire plunger tests at high rates and high temperatures. He
used a test machine which consisted of a pneumatic gun for propelling the
plunger at an inflated tire, mounted in a temperature-controlled cabinet,
and photoelectric devices for measuring plunger velocity. The minimum
kinetic energy required to cause failure of the tire fabric is taken as a mea
sure of carcass breaking energy. Data on rayon and nylon cord tires in
dicate that the breaking energy of a tire is dependent on both speed and
temperature and that the relationship between energy, speed, and temper
ature depends on the cord-reinforcing material. As a result, it is impossible
to predict the relative impact resistance of tires under service conditions
on the basis of laboratory test results obtained with the standard static
plunger test, in which the plunger penetrates the tire at 2 in/min. at room
temperature.
A comparison of data obtained on 7.75-14, two-ply rayon and nylon
cord tires at three different speeds is shown in figure 2.27.
O'Neil, Dague, and Kimmel [120] also reported test results which in
clude results of an individual cord impact test and three dynamic tire tests
using a pendulum, a ballistic plunger, and resiliometer bruise.
Figure 2.28 shows the stress-strain curves of the four cord materials
used.
Effects of temperature on strength and breaking energy of these cords at
low speed are shown in figure 2.29a and 2.29b respectively. Decreases in
strength and breaking energy of nylon are more rapid than those of rayon
and polyester.
Effects of temperature on strength and breaking energy of these cords at
high speed are shown in figures 2.29c and 2.29d, respectively. Also in this
case, decreases in strength and breaking energy of nylon are greater than
TIRE CORD AND CORD TO RUBBER BONDING 71

oooooo

8S3
i i T
o-
— i

I I I

.3X10 sec-'
O-
oo o O oo o
X '«
^ *

(N 1 c*1
a 1 1
oo o\ V r~ oo v
06 06 oo -- «n en
UJ IN <*> ^ >o ^r oo

^r <N o t «n o + +

X 'C —• r 3- c

27 •X 0- — r ( ^_ TT
07
^^ I I I

poo —
Ov ON
— (N
i i i

Material 11 -all
fiiliS
IHIil
«nti
•3-3gghH
55^-SHH
72 MECHANICS OF PNEUMATIC TIRES

— M"" fNCN fN<N fN — "-

**"* "" °° °°
u. ™,

fNfN^fNCS^fN — ""fS — ^

O^or-ift--ftftOfN^i: — m
CS fS ft fS ft O< ft fN

— O»N»»ie><N3OH

JS-P
•p •g •g

8
IM
3 3
O *o *o
.9 B .S

.o.S'j.o.S" o.S-o.S
rjf

£ o S KI S
r-
s g
"

_o co-

2HI!
TIRE CORD AND CORD TO RUBBER BONDING 73

TESTING SPEE 9
— •RAYON 2IN/MIN
—T
• 63MPH
(INCH
BREAKING
ENERGY
POUND) 100 MPH
— o NYLON 2IN/MIN
—7 63 MPH '

t
—a 100MPH
x
I
N^ ^
^ O^ __

I r

"^ ^__
^
»-_ *

'—
L _\ )

75 125 175 225


TEMPERATURE^ F )

FIGURE 2.27 Breaking energy of 7. 75-14, two-ply, rayon and nylon cord tires as afunction of
temperature.
Speed: 2 m/nun . 63 mph. and 100 mph.

those of rayon and polyester. On the other hand, rayon shows an increase
in breaking energy with temperature.
They also examined the impact resistance of tires. Table 2.24 shows the
comparison between the low speed test and the high speed tests. The nylon
cord tire has the highest failure energy at low speed, and the polyester
cord tire has the highest value at high speed.
2.1.12. Fatigue Resistance of Tire Cords
Fatigue resistance of tire cord is an important property but difficult to
assess. A variety of cord fatigue tests and laboratory tire wheel tests are
currently available for evaluating relative tire durability. However, none
of these tests can be considered adequate for characterization of road per
formance and therefore, in the last analysis, a road durability test is re-
auired.

MYLDKI

FIGURE 2.28 Stress-strain curve, low speed, 12 in/min. [120]


74 MECHANICS OF PNEUMATIC TIRES

MO 400 100 200


TEMPERATURE ( *F )

FIGURE 2.29 of temperature on breaking strength and energy [120]


A and B Low speed test C and D High speed toL

Numerous papers are concerned with fatigue resistance of tire cord;


some give phenomenological treatments of fatigue under cyclic tension
[122-125] biaxial rotation [126] or flexing of cords in air [123, 127-131],
some present statistical treatments of fitigue phenomena [132-184] [130d]
and others are treatments of fatigue under compressive or flexing condi
tions in rubber blocks or in actual tires.
Here, we will be mainly concerned with fatigue in rubber blocks or in
the tire itself, which is most important for durability of tires.
Various views on the mechanism of cord fatigue in tires have been pre
sented in the literature and diverse opinions have clearly indicated need
for further research. For example, Williams and co-workers (135) con
cluded that cords in tires lose their strength linearly with mileage due to
broken filaments. On the other hand, Entwistle and co-workers, and Klein
and co-workers (136), see figure 2.30, said (137) that cords did not become
progressively weaker until failure but fail suddenly by an undefined cata
strophic process.
Fatigue failure of cords in tires generally occurs at special localized
points. Williams and co-workers [135] and Patterson and Anderson re-

TABLE 2.24 Plunger, Pendulum, and projectile tests of tires [120]


rayon nylon polyester
Plunger (slow) (I)
energy, in-lb 2260 3450 3120
Pendulum (impact)
energy, in-lb 9160 3390 3900
change over (I), % +40 -2 +25
Projectile (impact)
energy, in-lb 3470 3600 4430
change over (I), % -1-53 +4 +42
TIRE CORD AND CORD TO RUBBER BONDING 75
• FLEX
STRCORD
ENGTH(LB) H CENTER
IMrsiMuOf-
K|
Co —p SIDE WA LL-
t
\ I

RAYC H
T-
1650/2
1 s
£20
1
20 40 60 80 100
MILES

FIGURE 2.30 Cord strength as a Junction of tire miles, taxi fleet test [136]

ported that cords lose their strength more rapidly in inner plies than in
outer plies. Klein and co-workers [136] also reported from tests of Tyrex
rayon and nylon cord tires, that cord strength loss is greater in the flexing
or side wall region than in the crown or center region, figure 2.30.
Patterson and Anderson [137] found that strength loss of nylon cords in
tires was affected by the direction of tire rotation and the direction of cord
bias and that cords in opposite sidewalls of tires lost strength at markedly
different rates. Strength loss was higher for the half of the cord that led
into the load bearing region of tire as it rotated (leading half) than for the
other half of the cord (trailing half), as shown in figures 2.3 1 and 2.32
[137].
Two regions of high strength loss were found in each tire sidewall by
breaking short segments of cords; at the shoulder, and at a point about two
inches above each bead. Figure 2.33 shows breaking strength at these
points in first-ply cords from unf ailed tires as a function of miles run on
the test wheel. The point of lowest cord strength, in the range from 12,000
to 18,000 miles where tire failures began to occur, is shown to be in the
leading shoulder. However, tire failures were always in the sidewall about
two inches above the bead. This seemed to indicate that tires do not fail
where the cords are weakest.
Patterson [138] concluded from the above discussions and microscope
examination of broken ends of cords, that fatigue failures in nylon fila
ments were mainly related to cyclic flexing associated with compressive
loading of inner ply cords.
In addition to the above fatigue failure of cords, it is generally accepted
that cord-to-rubber adhesion failure [136] and rubber failure [139] play an
important part in fatigue failure of tires. The general levels of adhesion (as

TRAILING HALF OF CORD

AXLE SIDE

URB SIDE

LEADI NG HALF OF CORD

FIGURE 2.31 Identification of cord sections to side wall and rotation direction (137)
MECHANICS OF PNEUMATIC TIRES

TRAILING HALF
STRENGTH
(t LEADING HALF
K
iv
CO
o ^—
v\ **. ——. ~—
^s.
^
i •*,
^ -—. "= • !•.

* ~-
»._ ^- *"fc
""»
«», -^

10
PLY 1 **.
PLY 3
1 A
WOO 2000 3000 0 XXX) 2000 30(
MILEAGE

FIGURE 2.32 Cord Strength as a function of tire miles at test [137]

measured by the stripping test) before and after a long run are shown in
figure 2.66. Highest adhesion loss is observed in the region of maximum
compressive flexing.
Microscope examination of broken ends of cords from failed nylon 66
tires showed a large number of filaments broken at a certain angle to the
fiber axis [137, 138]. Polyester and nylon 6 filaments showed similar bias
ruptures [137, 140, 142].
Bias breaks are most prevalent at the point of cord rupture but are also
found occasionally along the entire cord length. Patterson and Anderson
stated that the number of bias ruptured filaments in flex tested tires in
creased linearly with severity of fatigue damage to the cord, see figure 2.34
[137]. These bias breaks were not produced by tensile loading of fila
ments.
Patterson dealt with the mechanism of bias rupture in another paper, a
number of tests being run on filaments and cords to produce bias rupture.
A variety of tensile loading conditions applied to single filaments failed
to produce bias ruptures. These results indicated that possibly unique or
complex loading conditions were imposed on cords during flexing in a tire.
To establish what these conditions might be, an investigation was made of

( ENGTH •>oi»o
KSTRG) «jM— K . |
T Rt IL ING
1\ B•-. iAD-
ht-
\\ \
TF AILI MC
\ \ s SHOU 1C ER
V k-
:A > 5g~~*1
E £/ . ^> -
^
LEA bi NG < HOULDER
|
4 6 12 16
MILEAGE

FIGURE 2.33 Average breaking strength of sections offirst ply cords [137]
TIRE CORD AND CORD TO RUBBER BONDING 77

STRCORD
EN&THtKG 3
U
O
01

N\
\
^v
K

PERCENT OF Fl LAMENTS
WITH BIAS BREAKS

FIGURE 2.34 Cord strength as a Junction ofpercent of bias filament breaks [137]

the behavior of cords in transparent rubber under a variety of loading cir


cumstances.
Examination of the cords during alternate tension and compression
showed that the filaments undergo bending and sometimes buckling when
in compression. The seventy of bending increased with reduction of twist.
Similar results were also obtained by Wood and Redmond [143]. The ob
served increase in filament bending at low twist and the accompanying
poorer fatigue resistance indicated that bias filament ruptures might not
depend on the complex stress associated with a twisted structure but more
directly on simple bending associated with compressive loading of the
specimens.
This hypothesis was tested with Mallory and Goodrich Disk Fatigue
Testers. These tests established conclusively that bias rupture was pro
duced by simple bending.
Further investigation showed that bias rupture began at the compres
sion side of the bend in filaments subjected to repeated cyclic bending.
Similar findings were indicated for ductile metal.
Fujimoto [42] investigated the correlation between fatigue life and mo
lecular structure of nylon 6. He found that fatigue life in compressive
loading decreased linearly with the second moment of the NMR peak of
the noncrystal line part, and then concluded that long fatigue life is to be
attributed to weak intermolecular interaction and mobility of molecules in
amorphous regions, see figure 2.35a. Furthermore, he mentioned that fa
tigue life decreased linearly with lattice spacing d (002), see figure 2.35b
and suggested that mechanisms of fatigue failure and strength loss in the
curing process, as mentioned before, have something in common.
Each tire company has its own methods of assessing tire cord fatigue re
sistance. Laboratory machines designed to examine fatigue resistance of
cords fall into two groups; those flexing cords with air as medium and
those flexing cords in rubber. Among typical examples of the former, there
are cyclic tension types (e.g., U.S. Rubber Fatigue Tester, Du Pont Dy
namically Balanced Tester [125] and Goodrich Tension Vibrator [124]
and flexing types, e.g. Goodrich Flexing Tester [128].
78 MECHANICS OF PNEUMATIC TIRES

,
X
x N^ ^

N, V
NX X
A B

1* IS !.» 1.0 13 TU 3.X IJt 311 3tt>


<iH2>A GAUSS2 d(002)A

FIGURE 2.35 (A) Fatigue life of nylon 6 as a function of second moment ofNMR peak for
amorphous region. (B) Fatigue life of nylon 6 as a function of lattice spacing d(002) (42).

Dillon classified these test methods according to the types of applied


stress or strain as figure 2.36 shows.
Of the latter group, generally regarded as the more important, there are
several methods, as we will mention below; Mallory Tube [145-147], U.S.
Rubber Tube [146], Goodrich Disk [146-149] Firestone Compression
Flex. [146-147] Dunlop and De Mattia Flex [150-153].
(1) Mallory Tube Test
The testpiece is a rubber hollow cylinder in which the cords to be tested
run parallel to each other and to the axis of the test cylinder and are ar
ranged to have the required number per inch.
The flexing principle is illustrated in figure 2.37. The test tube is bent
and tightly clamped on two spindles of the flexer. Thereafter, air pressure
inside the tube is increased and then the horizontal spindle is rotated.
Cords in the tube undergo alternatively compression and tension.

C\AA/V\A7
(A)"1 (D)

(B)
I

(F)

TIME TIME

FIGURE 2.36 Classification of cyclic tension fatigue tests (144).


TIRE CORD AND CORD TO RUBBER BONDING

DRIVE
PULLY
COUNTER

REGULATED AIR
.PRESSURE SUPPLY

FIGURE 2.37 Schematic drawing of Mallory Tube Fatigue Tester.

The number of revolutions until failure is the measure of fatigue resis


tance.
(2) U.S. Rubber Tube Test
This method is basically similar to the Mallory Tube Test but the flex
ing principle differs somewhat, as shown in figure 2.38.
(3) Goodrich Disk Test
The flexing principle is illustrated in figure 2.39. The testpieces are rec
tangular rubber blocks in which test cords run parallel to the long axis.
The testpieces are firmly secured into the periphery of two canted disks,
the cords passing across the gap between them. When these disks are motor
driven, each cord in a rubber block sutlers simple longitudinal extension
and compression. Strength loss after a certain number of revolutions is ex
amined to evaluate the cords.
(4) Firestone Compression Flex Test
The testpiece is a rubberized belt which contains two plies of cords in
parallel planes, one ply of test cords and the other of steel cords.
Compressive fatigue is produced by flexing the testpiece over a spindle,
figure 2.40. The inextensibility of the steel ply causes the layer of rubber
itself and the spindle to be compressed and the tire cords are situated in
this layer.

ROTATING ARM

^ROTARY UNION

CLAMP

REGULATE AIR
PRESSURE SUPPLY

FIGURE 2.38 Schematic drawing of U.S. Rubber Tube Fatigue Tester.


80 MECHANICS OF PNEUMATIC TIRES

SPECIMEN

ROTATING SHAFT
FIXED

SPECIMEN

FIGURE 2.39 Schematic drawing of Goodrich Disk Fatigue Tester.

(5) Dunlop Fatigue Test


Essentially, the test piece is an endless belt made up of five plies of rub
berized cord. Counting from the inside, the second and fifth plies are com
prised of the cords to be tested, these cords running along the length of the
endless belt and being arranged in the required number per inch. The belt
is tensioned between two pulleys as shown in figure 2.41. The test usually
consists of running belts for known times, then extracting the cords from
the two test plies and measuring their breaking strength.
(6) De Mattia Flex Test
The testpiece is a rubber block, in which the cords to be tested run par
allel to each other and to the long axis of the block. The testpiece is firmly
clamped on the two heads of a De Mattia Flex Tester which is usually
used in rubber flex fatigue tests, figure 2.42. The test consists of flexing the
rubber blocks and measuring their residual strength.

2.1.13. Effects of Twist on Cord Properties


Many papers have been published dealing with e ffects of twist on cord
properties, some are theoretical, others experimental.

SPECIMEN

PIVOT

FIGURE 2.40 Schematic drawing of Firestone Compression Fatigue Tester.


TIRE CORD AND CORD TO RUBBER BONDING 81

•DRIVEN PULLEY

TEST BELT

- IDLER

J APPLIED LOAD

FIGURE 2.41 Schematic drawing of Dunlop Fatigue Tester.

Tire and tire cord producers throughout the world do not agree on
which twist gives the best results. Some of them choose a symmetrical
twist and others prefer an asymmetrical twist.
They also adopt different twists for different materials; usually polyester
cords are twisted more highly than nylon cord. Generally speaking, as
twist increases, cord strength, initial modulus, and fatigue resistance in cy
clic tension decrease, while elongation at break, rupture energy, and fa
tigue resistance in compression increase.
Kemmnitz and co-workers have reported extensive studies on rayon
cord dealing with effects of twist on cord properties. Figures 2.43 to 2.50
[154] show their experimental results. Cord construction was HOOD/2.
The effects of twist on strength were observed when the ratio of ply and
cord twist was not equal to unity. Cord strength is indicated as a function
of twist shrinkage in figure 2.44.
In figure 2.45 rupture energy is shown as a function of twist shrinkage.
Figures 2.46 and 2.47 show the relationship between dynamic properties
and twist shrinkage.
These cord properties were stated to be uniquely determined by a
simple function of twist shrinkage in the range examined. In figure 2.48 ef-

CLAMP

FIGURE 2.42 Schematic drawing of De Mania Flex Fatigue Tester.


82 MECHANICS OF PNEUMATIC TIRES

397 ( TURNS/ IOCM)

0 20 40 60 80 100 120
SINGLE TWIST (TURNS/IOCM)

FIGURE 2.43 Effect of twist construction on breaking strength of raw cords [154]

? II
s" \
QC
\
m *
i* ~ s\ s
V
\
\s
9 3
i
0 5 10 15 20 25 30
TWIST SHRINKAGE (V. )

FIGURE 2.44 Breaking strength of cord as a function of twist shrinkage [154]

ENIMPACT
RUPTURE
ERGY(C
CB
*O»
O
*SJ* \
Ml
W
Wl
\
\
\
\
\
1 J
\
i
\
0 10 20
TWIST SHRINKAGE (•/, )

FIGURE 2.45 Impact rupture energy as a function of twist shrinkage [154]


TIRE CORD AND CORD TO RUBBER BONDING

12

* 8 12 16 20 24
TWIST SHRINKAGE V. )

FIGURE 2.46 Energy damping as a function of twist shrinkage [154]

XIO*

in
Sw

uiz
0.8

<. 8 12 16 20 12
TWIST SHRINKAGEC/. )

FIGURE 2.47 Energy loss as a junction of twist shrinkage [154]

PLY TWIST
12 NUMBER 50TURNS/10CM
o OF CYCLES ' 59
X 70
-1
i

UJ
K

_in ^
UJ
oe
0 tO SO 120
SINGLE TWIST (TURNS/ 10CM)

FIGURE 2.48 Residual strength after 105 and Iff cycles offlexing (De Mania Tester) [154]
MECHANICS OF PNEUMATIC TIRES

OF
CYCLES
FLEXING
(
•F
A /.

S10
AFTER
LOSS
TRENGTH |
S
gD J
/
'
/

/
/
/
/
/
/20 10 5 ZV
01 1 02
ZV
TWIST LOSSC/.)

FIGURE 2.49 Correlation between strength loss after Iff cycle offlexing and twist loss [154]

fects of twist construction on fatigue resistance are shown as residual cord


strength measured after a known period of flexing on the De Mattia Flex
Tester.
The relationship between residual strength after a known period of flex
ing, ∆F , and twist loss of cord ZV (Zwirn Verlust) are indicated in figure
2.49.
∆F decreases linearly with ZV and the optimum residual strength is ob
tained by suitably balancing ∆F with ZV.
Figure 2.50 also shows the relationship between amount of twist and
cord strength after flexing with number of load cycles as a parameter
[155]. Twist construction was a symmetrical twist of 1650D/2. Obviously
the symmetrical twist 49 turns per 10 cm. reached its peak strength already
after application of few load cycles. Similar results have also been ob
tained with Disk Fatigue Tests and tire fleet tests [135] figure 2.51.
Furthermore, it is well known that fatigue resistance measured by the

NUMBER OF
1,3 CYCLES
0

ac
5 H
2 10

40 48 51 60
TWIST (TURNS/ tOCM)

FIGURE 2.50 Residual strength of cord afterflexing on De Mania Tester as aJunction ofply
twist and cycles offlexing [154]
TIRE CORD AND CORD TO RUBBER BONDING

STREN&TH
NG
BREAK!
B! :FO X TIRE FLEET TEST
- TIS
f~
O
CO
•>
*J
_F .EX
... -DIS TEST
™"-»^,
--. ^^ —
>
^s ^
*^s AFTER F LEX
/
* 10 II 12 13 14 15 16
TWIST (TURNS /INCH)

FIGURE 2.51 Effect of twist on cord strength before and after flexing in actual tire and
Goodrich Disk test [135]

Mallory Tube Test also increases with amount of twist in the ordinary
twist range [68, 69).
Fujimoto stated that fatigue life in cyclic tension decreases with the
amount of twist in the ordinary twist range, while flexing and compressive
fatigue life increase, figure 2.52.

2.2. Rubber-to-Cord Bonding


2.2.1. Introduction [83]
There have been many reports describing the various points of view on
the adhesion of cord to rubber. This chapter summarizes the authors'
views on the subject. An attempt has also been made to determine from
both patents and the literature how the adhesion of cord to rubber is being
improved.
The adhesion of cord to rubber differs greatly from that in any other
field. The properties required for cord-to-rubber adhesives are summa
rized as follows:
1) There are significant differences between fiber and rubber in modu
lus, molecular polarity, and reactivity. For the adhesion of fiber to rubber,

LIFE(MIN)
FATIGUE i- u
§
8
5 ^
E* fErs SIC)N
f— ^ ,'' M ALL ORY TUBE
^ ^ > ( POl V ESTER )
,'*
j
^ \
v
&
16 50/ 2 MY ON
%
\
2 A 6 8 10 12 U 16 18
TWIST (TURNS/ INCH)

FIGURE 2.52 Fatigue life of cord as a function of twist [156]


86 MECHANICS OF PNEUMATIC TIRES

it is desirable that the adhesive have an intermediate modulus and two-


faceted properties in terms of molecular polarity and reactivity. Therefore,
rubber blended with polar and reactive resin is widely used in practice.
2) Cord-rubber composites require flexibility in most cases, so flexible
adhesives with a rubbery matrix are desirable.
3) Heat and fatigue resistance are required. To meet the requirement, it
is necessary for an adhesive to have a three-dimensional network struc
ture, to be nonfluid under high temperature and stress, and to have a high
tenacity.
4) Water-base adhesives are required for operational reasons.
5) Resin ingredients are fluid in the baking process and must change to
an insoluble polymer when baked.
2.2.2. Outline of Bonding Methods [83]
Adhesives for rayon and nylon:
RFL, made by the addition of resorcinol-formaldehyde condensate to
rubber latex, is popular as an adhesive for nylon and rayon. To obtain the
best adhesion from the RFL treatment, such factors as recipes, maturing
conditions, and baking conditions including squeezing and drying condi
tions must be optimum. These conditions affect not only the adhesion but
also the appearance of dipped cord fabrics (blister), machine cleanliness,
liquid viscosity, coagulation stability, foaming, and cord strength. Thus,
they require attention when the RFL treatment is used.
Consideration is now given to the role of each component in the RFL.
Latex components make the layer of an adhesive flexible and combine it
firmly with the rubber layer by means of the secondary bond and co-vul
canization. The ratio of SBR latex to VP latex is determined by the rubber
compound to be adhered (for example, the blending ratio between SBR
and NR). From our experience, the ratio of SBR latex to VP latex which is
likely to produce stable adhesion ranges from 20/80 to 30/70.
RF components, however, react to the active hydrogen of the OH group
in rayon and of the NHCO bond in nylon, forming the primary bond or
the hydrogen bond with the fiber. This also forms a three-dimensional net
work structure. Thus, they are thought to reinforce the layer of an adhe
sive. According to the report by Patterson, only an adhesion is formed by
the hydrogen bond if RF components are simply dried after dipping. This
adhesion is easily spoiled by the intervention of water molecules, and is
weak in water resistance; this reaction is reversible. However, a water-re
sistant, strong primary bond is formed when the components are baked at
high temperature. [ 1 57]
There has also been the reaction mechanism in which RF components
react to a double bond or react to the a-methylene group of rubber mole
cules, forming the primary combination. However, it is not clear how im
portant a role the reaction mechanism plays [158-160].
Figure 2.53 shows interactions of rubber-adhesive (RFL)-fiber, as de
scribed above.
Adhesives for polyester:
A normal RFL treatment does not provide sufficient adhesion for poly
ester cord to rubber. This is thought to be mainly because there is no ac
tive site other than the terminal OH and COOH group on the polyester
surface.
TIRE CORD AND CORD TO RUBBER BONDING 87

RUBBER

»• LATEX

FIBER

FIGURE 2.53 Interaction of Rubber-Adhesive-Fiber

For this property of polyester fiber, new adhesion methods have been
developed by many researchers. They are broadly classified into the fol
lowing:
1) Single-dip method with the use of isocyanate rubber cement (rubber
cement method).
2) Single-dip method using a mixture of very reactive low molecular com
pounds such as phenol blocked isocyanate, isocyanate dimer and ethyl-
ene urea, and rubber latex or a RFL (reactant type single-dip adhesion
method).
3) Single-dip method in which the resin components of the RFL are re
placed by phenolics which have high affinity to polyester. For example,
Toray's "1" and Id's "Pexul" (Adsorption type single-dip adhesion
method) [38, 161, 162].
4) Double-dip or raw yarn treatment method in which the fiber surface is
pretreated with an epoxy compound, isocyanate, polyethylene-inline or
modified PVC and then is RFL-treated (multistage adhesion method).
Of these 4 adhesion methods, 1 and 2 are of little real use due to the non-
aqueous system in the case of 1 and the decline in adhesion when vulcan
ized at high temperatures as well as the settling stability of dispersion sys
tem in the case of 2. Therefore, 3 and 4 are those most widely used.
Method 4, epoxy/RFL double-dip method and its modifications are par
ticularly in frequent use.
Addition of bonding agents to rubber compound:
Recent research has been done on new adhesion methods. In these
methods the adhesive components are mixed with the rubber for the direct
adhesion of cord to rubber without any prior RFL treatment of the fiber.
A typical example of the adhesives used is an isocyanate derivative and a
combination of a methylene accepter and resorcinol, with a methylene-do-
nor such as hexamethylene-tetramine.
These new methods eliminate the dipping process completely for nylon
and rayon fibers, and require only a pre-treatment with epoxy for polyes
ter to obtain satisfactory adhesion.

2.23. Adhesion Mechanism [83]


There have been a good number of reports on the adhesion mechanism
of tire cord. However, they are not complete and are generally still in the
88 MECHANICS OF PNEUMATIC TIRES

stage of speculation when considering actual problems. This field has


made "technology-ahead" progress and all the new techniques can be said
to have been established only after much groping.
Mechanical adhesion:
As has been well known from early times, the anchoring effect of fluff
on the fiber surface and bridging effect of fabrics produce satisfactory re
sults in the adhesion of fiber to rubber. For tire cord, the RFL penetrates
between filaments to some extent and some anchoring effects can also be
expected. In order to control the penetration of the RFL into the cord, it is
necessary to make approaches from three points, i.e., RFL recipe, dipping
conditions and spin finishes of yarn.
Adhesion interface and place of bond rupture:
In terms of adhesion systems, the breaking force is usually called adhe
sion. However, rupture does not always occur on the interface. In the case
of cohesive failure, the breaking force does not represent the adhesion on
the interface.
The force in the pull-through and peel tests is affected by the modulus
of elasticity of the rubber to he adhered. In addition, locations of rupture
are changed from the interface to the cohesive ruptures depending on the
tension speed and thickness of interply rubber. These facts are well
known.
To simplify such a complicated phenomenon, let us consider the cross
sectional model of the adhesion interface, as illustrated in figure 2.54.
The first step necessary for research on adhesion is to observe rupture
occurring during an adhesion-breaking test. The observation of where the
rupture occurs is very difficult, though it might seem to be easy. It is par
ticularly difficult to distinguish the rupture in layer (3) from that in layer
(4). In the case of RFL-treated nylon or rayon, rupture often occurs in
layer (1) and at the interface (2), requiring no close attention to the rup
ture on any other layer.
As far as polyester is concerned, the adhesion to the cord surface is diffi
cult causing the rupture to occur at the surface (4). In addition, the fiber
surface deteriorates under action of the low molecular uniine compounds
and the moisture contained in rubber, resulting in a rupture at layer (5)

Rubber
o
Adhesive

Cord

FIGURE 2.54 Adhesion interfaces


TIRE CORD AND CORD TO RUBBER BONDING 89

(appearing as a single cut filament of a cord cut). As previously described,


specially compounded adhesives are often utilized to increase the adhe
sion on the surface of polyester. Some of these adhesives are insufficient in
the strength of their layer, possibly resulting in a rupture of layer (3) in an
adhesion test under high temperatures. In fact, there are many cases in
which ruptures of layers (1) through (5) occur in a mixed state. However,
this mixing ratio cannot be ascertained quantitatively. This makes the re
search more complicated.
Bond between cord and adhesive interface:
The following factors are considered to contribute to the bond of a cord
with the adhesive surface:
1) Formation of the primary bonds.
2) Formation of the hydrogen bonds.
3) Dispersion force effect (related to the solubility parameter).
4) Diffusion of adhesive components into the internal structures of the fi
ber molecules.
5) Fluid process of adhesives on the fiber surface.
When treating rayon or nylon with RFL, factors 3 through 5 require
little consideration because factors 1 and 2 make the major contribution to
the adhesion.
In the case of polyester, however, factors 3 through 5 are important be
cause factors 1 and 2 play only a small role. The formation of the primary
bonds depends on the reaction of the terminal OH group, the terminal
COOH group and the COO bond. For an epoxy adhesive, anything but
the reaction of terminal COOH group can be ignored. The quantity of the
COOH group has no effect in the case of isocyanate or "1".
The molecules of polyethylene amine contain primary, secondary and
tertiary amines, which act on the COO bond of the polyester to facilitate
the formation of adhesion.
Factors 2 through 4 are important for the case of "I" or "Pexul".
Iyengar and Erickson have substantiated that excellent bonding of poly
ester film could be obtained with the use of an adhesive whose solubility
parameter was near 10.3 of that of polyester, figure 2.55. They also say the
Du Pont's epoxy-isocyanate adhesive is based on this idea. [163].
Id's "Pexul" seems to follow the same idea to a certain extent.
Now, we will consider the diffusing process of an adhesive into the
structures of the fiber molecules.
On analysis by GPC, it has been shown that "I" or "Pexul" contains in
its components low-curing speed and low-molecular weight phenolics
which have no methylol group. These phenolics diffuse within the fiber's
molecular structure, forming a bonding site on the fiber surface. In other
words, the adhesion decreases when resin components are replaced with
many methylol groups containing high-polarity resole type components,
which contain many methylol groups and have a high curing speed, or
when the molecular weight is increased.
Adhesive-activated type polyester is another example in which the dif
fusion of the adhesive into the fiber's molecular structure is very impor
tant. When yarns are drawn after applying an adhesive to amorphous un
drawn yarns the adhesion is higher and more efficient than that obtained
when applying the adhesive to drawn yarns.
The fluid process of adhesives relates to the application of an adhesive
90 MECHANICS OF PNEUMATIC TIRES

8 9 10 tl 12 13
SOLUBILITY PARAMETER OF ADHESIVE

FIGURE 2.55 Relationship between peeling strength ofpolyesterfilm and solubility parameter
of adhesive [163].

to the fiber surface, uniform coating and fluid viscosity. However, no de


tailed research has been made on this process.
Bond between adhesive and rubber interface:
The following factors are considered to affect the bond of an adhesive to
the rubber surface:
1) Covulcanization between latex components in the adhesive and the
rubber.
2) Dispersion force effect of latex components and the rubber.
3) Formation of a blended layer of entanglement of molecular chains
through the mutual diffusion at the surface of latex components and
the rubber.
4) Fluidity of the rubber layer during vulcanization.
5) Formation of the primary bond between adhesive resin components
and the rubber.
As mentioned earlier, the RF in the RFL may form the primary bond
with the rubber molecules. However, it cannot be considered as a main
factor controlling the bonding force of an adhesive with the rubber sur
face. This is because no bonding of the adhesive to the rubber surface re
sults from an incorrect combination of latex components in the adhesive.
This fact can rather be said to suggest the importance of the role which the
latex components play in the adhesion of the interface.
Factors 1 through 3 relate to the selection of type of latex components
for the rubber to be adhered. Similar to the ideas controlling cov-
ulcanization of a blend of different elastomers, the best result is obtained
when the rubber and the latex's solubility parameters and their vulcan
izing speed are matched.
The dispersion force and mutual diffusion effects are inseparable from
each other and the solubility parameter reasonably explains both effects.
The latter relates further to the mobility of rubber molecular chains. It is
TIRE CORD AND CORD TO RUBBER BONDING 91

affected by the vulcanizing temperature, curing of latex in RFL, Tg of


rubber, oxidation of the surface and development of a three-dimensional
network structure on the RF resin. Dietrick performed an experiment by
altering the copoly merization ratio of butadiene in the SBR latex, which
demonstrated that the greater the quantity of butadiene, the higher the ad
hesion. [164]. This phenomenon can be explained by any of factors 1
through 3.
As for the polarity of the rubber, consideration must be given not only
to the molecules of the elastomer themselves but also to the polarity of the
filler.
The connection between fluidity in rubber vulcanization and adhesion
is made clear by Iyengar's work. He experimented on rubber compounds
by using a variety of vulcanization accelerators, proving that the longer
the Mooney scorch time, the greater the adhesion, as indicated in figure
2.56 [165].
Our experimental results, however, show the effects of vulcanization ac
celerators to be very complicated and it is unlikely that high adhesion de
pends only on the scorch time.
The change in adhesion with aging appears in the surface of the adhe
sive on the cord. The effect of ultraviolet rays, heat, oxygen and ozone are
now under study as possible causes of the change. When maintained for a
long period of time, the dipped cord's color fades and its adhesion declines
at the same time. This phenomenon seems to be related to the change in
the RF components.
Rupture offiber layer:
The rupture of fiber layers is a difficult phenomenon to observe. During
an adhesion test, the cord surface sometimes fluffs because of a single bro
ken yarn, causing the cord strength and adhesion to decline. This is an ob-

10 15 20
SCORCH TIME.min

FIGURE 2.56. Relationship between adhesion and scorch time.


92 MECHANICS OF PNEUMATIC TIRES

vious example of the rupture in fiber layers. This phenomenon is observed


from time to time in the adhesion of polyester fiber. When compounding
special vulcanization accelerators such as thiuram, thiocarbamate and
hexamethylene-tetramine with the rubber, the fiber strength decreases and
ruptures in fiber layers occur, preventing efficient adhesion. [138].
Furthermore, when the composite of polyester fiber and rubber is ex
posed to a high-temperature atmosphere for a long period, the cord is
found to decline in strength and adhesion. This is thought to be due to the
deterioration of the polyester caused by the amines compounds and mois
ture in rubber. This phenomenon is a very important problem connected
with adhesion durability of a running tire. [79].
Rupture of rubber layer:
If a rupture occurs in the rubber layer during a peel test, the adhesion,
generally, is judged to be satisfactory. In the peel test, a rupture tends to
occur in the rubber layer when the rubber layer between plies is thick,
while if it is thin the rupture gradually shifts to the interface. So this phe
nomenon requires further attention. Detailed research has been made on
the phenomenon by peel tests of pressure sensitive adhesive tape. [166].

2.2.4. Adhesive Treatment of Nylon and Rayon-RFL Treatment


Many factors are known to affect the bond strength secured with RFL
adhesives. For example, composition of RFL, method of RFL prepara
tion, amounts of adhesive applied to cords, its distribution on the cord,
heat treatment after dipping, method of storing dipped cord, composition
of rubber compound, and vulcanization conditions may be factors affect
ing the cord to rubber bonding. Participation of each factor is explained in
the following sections.
RFL adhesives are usually prepared by reacting resorcinol and formal
dehyde under alkaline conditions prior to addition to latex. The mixture
of RFL resin solution and rubber latex is further aged before use. An ex
ample of RFL preparation is as follows [167].
Sodium hydroxide, resorcinol, and formaldehyde are dissolved succes
sively in water. After complete solution is effected, a reaction takes place
which should be allowed to continue for six hours at 25 °C.
Resin solution
Water 238.4 g
Resorcinol 1 1 .0
Formalin, 37% 16.2
Sodium hydroxide 0.3
The RFL recipe should be varied with the textile material. The follow
ing formulations have been found to give optimum adhesion with rayon
and nylon.
Fabric Rayon Nylon
Gentac latex, 41% 52.5 g 428.0 g
SBR latex, 40% 215.0
RF solution, 6.5% 284.0 465.0
Water 467.8 107.0
The recommended maturing condition for the above RF-latex mixtures
is six hours at 25°C.
TIRE CORD AND CORD TO RUBBER BONDING 93

Precondensation of RF resin is, however, not indispensable when syn


thetic latex is used [277].
Also resorcinol-formaldehyde precondensate, such as Penacolite resin
of Koppers Co., is available to prepare RFL [164]. Adhesive is prepared
by successively mixing the following components:
Water 407.7 g
Sodium hydroxide, 10% aq. sol. 8.0
Penacolite resin R2 1 70, 75% 26.7
Formalin, 37% 20.3
Vinylpyridine latex, 40% 250.0
The mixture should be aged 18 to 20 hours before use.
Adhesion is affected by the composition of RFL and maturing condi
tions. Resorcinol to formaldehyde ratio, pH of solution, concentration,
maturing time and temperature have an effect on the structure and molec
ular weight of resorcinol formaldehyde resin. Choice of latex component
is very important on the viewpoint of affinity of RFL for rubber. Latex to
resin ratio is also influential in adhesion. Strength of RFL film and bal
ance of affinity to both fiber and rubber depend on L/RF resin ratio.
Nowadays, the composition of RFL for nylon and rayon has been de
cided empirically. R/F is in a range of Vi to 1A in molar ratio, and L/RF
resin is in ' . to 1/,., in weight ratio of solid components. The most widely
accepted latex component for natural rubber and/or SBR compound is vi
nyl pyridine-styrene-butadeiene terpolymer latex. In some cases, mixture
of vinyl pyridine latex with SBR latex on natural rubber latex is used [34,
277, 164, 168-186].
Adhesion is affected by conditions of dipping and of heat treatment af
ter dipping.
Effect tuid control of RFL pickup: Pickup of adhesive on a cord and pen
etration conditions of the adhesive into the cord interior affect adhesion.
From general experience, bonding strength increases with pickup of adhe
sive. It is, however, important to consider not total pickup, but effective
pickup since penetrated RFL does not contribute to adhesion. It is ob
served by microscopic inspection of a cross section that RFL penetrates
into the cord interior to a certain extent.
Penetration may be controlled by squeezing conditions [ 1 54], concentra
tion of RFL, viscosity of RFL, cord tension in the dipping bath [187], and
presence of water predip. However, the contribution of penetration of
RFL to adhesion has never been quantitatively analyzed.
The dependence of adhesion on RFL pick-up was reported in many pa
pers.
Adhesion increases gradually with pickup and reaches a saturation
point. An example is shown from Dietrick [164], and illustrated in figure
2.57. Too much pickup should be avoided because the amount of adhe
sive affects stiffness of the cord [188]. Also a pickup level should be de
cided upon by balancing cost against adhesion level. It is generally recog
nized that an adequate level of RFL pickup is from 6 to 8 percent of the
cord weight.
Pickup level is affected by the kind of textile material, textile lubricant,
concentration and viscosity of adhesive solution, and conditions of squeez
ing. Looking at the kind of textile material, nylon takes up the adhesive
with more difficulty than rayon because of its hydrophobic nature. Polyes
ter has still less pickup than nylon. A comparison for rayon and nylon is
94 MECHANICS OF PNEUMATIC TIRES


)
""" 9 *
5 '* )— O

3 10

' !L.
02466
ADHESIVE PICK-UP (%)

FIGURE 2.57. Effect of adhesive pick-up on pull through load [164].

shown in table 2.25 [277]. If dipping is done under the same conditions
where concentration of RFL and squeeze pressure is held constant, rayon
cord has more RFL pickup, and adhesion is superior to that of nylon cord.
Concentration of adhesive should be changed with the textile material to
adjust the pickup level. For example, it is recommended that suitable con
centrations for nylon and rayon are 20 and 12 percent respectively [167].
Even if the textile material is the same, different lubricants cause differ
ent pickups and different bond strengths. Gillman and Thoman [189] re
ported on this problem using the casein-latex adhesive system with rayon
tire cord. When effects of waxy ester and sulfonated oil are compared, hy-
drophilic cord lubricated with sulfonated oil has higher pickup than hy-
drophobic waxy ester cord. Materials that have been used in tire cord lu
bricants are white mineral oil, petroleum sulfonate, triglyceride,
ethoxylated and sulfonated derivatives, and so on. Generally, two or more
of these materials are blended to produce a proprietary formulation. Pe
troleum sulfonate is often used to enhance cord to rubber adhesion [190].
The dependence of adhesion on lubricants is, however, a very complicated
phenomenon, and cannot always be explained solely by the hydrophile-
hydrophobe nature of the lubricants.
Also, pickup level depends on the adhesive concentration and squeeze
conditions as shown in table 2.26. Change of concentration is more effec
tive than change of squeezing conditions.
Effect of heat treatment after dipping: Heat treatment conditions should
be decided from the kind of textile material, adhesive composition, and

TABLE 2.25. Relationship between squeeze conditions andpickup (277]


Rayon Nylon
Pull- Pull-
RFL Squeeze Pickup, through Pickup, through
Latex RFL cone., % roll, !h % load, Ib % load, Ib
Butyl latex 12.5 0 8.5 17.2 3.3 11.4
12.5 15 X2 14.5 2.8 10.7
12.5 35 3.5 12.5 2.2 10.6
SBR latex 0 73 264 3.2 177
15 43 23.9 2.9 15.9
35 3.7 19.2 2.5 14.2
TIRE CORD AND CORD TO RUBBER BONDING i

TABLE 2.26 Effect ofhydrophUic nature oflubricants on adhesive pickup and adhesion

Waxy ester on cord Sulfonated oil on cord


Viscosity
Adhesive at Pickup, Adhesion, Pickup, Adhesion,
cone., % 25°C,cp % Ib % Ib
5 1.67 0.4 7.8 1.7 9.0
7.5 1.94 0.5 8.0 2.6 10.6
10.0 2.46 0.8 8.1 3.9 10.9
12.5 3.53 1.2 9.4 6.1 12.2
15.0 8.36 1.9 10.9 8.4 13.8
17.5 26.5 2.5 11.0 9.5 14.1
20.0 57.2 5.9 12.0 22.0 17.4
22.5 83.5 21.1 14.8 46.3 19.8

procedure for the condensation reaction of the RF resin. Since each heat
treatment equipment has its own heat efficiency, it is impossible to estab
lish identical heat treatment conditions. Roughly saying, commonly ac
cepted conditions are:
Rayon 155° to 165° C 2to3min.
Nylon 6 205° to 210° O.Stolmin.
Nylon 66 220° to 230° 0.5 to 1 min.
During heat treatment, a highly crosslinked structure is formed in RFL
and strong interaction between adhesive and textile is achieved. Weak
heat treatment causes the RFL coating to have inferior tensile properties
and there is lack of interaction with the textile. Strong treatment impairs
compatibility with the rubber. Both over- and under-heat treatments are
undesirable for adhesion. These features are shown in figure 2.58. lyengar
reported the similar results on nylon 66 and 6, [191].
Recent improvement of RFL adhesives [83]:
Although RFL technology has been long established, there have been a
variety of improved methods proposed in recent years to meet the de-

~ 5
HEAT TREATMENT AT
160-C

0 2 4 6 8 10
HEAT TREATMENT TIME (MIN.)

FIGURE 2.58. Variation of adhesion with heat treatment conditions.


96 MECHANICS OF PNEUMATIC TIRES

mands for better adhesion under severe service conditions and for use with
diversified rubber compounds.
(1) Preventing deterioration of adhesion by aging
RFL-dipped cord will gradually deteriorate in adhesion by aging. It is
evident that light, air, heat and other factors are involved in the deteriora
tion, although the mechanism is not sufficiently clear, as stated previously.
To prevent light-induced deterioration, a pigment which absorbs light of
specific wave lengths is added [192]. To prevent latex components from
deteriorating through oxidation, double RFL treatments are applied with
increased latex contents for the second RFL. [193].
In addition, there is a method proposed to improve aging resistance by
adding an ampholytic surface active agent to the RFL. [194].
(2) Improvement of RF resin
The degree of condensation of the RF affects adhesion. The RF is usu
ally used after maturing at room temperature. However, there is also a
method in which pre-condensed resin is used. RF-precondensates using a
metal salt of carboxylic acid like zinc acetate as a catalyst can be used in
another method. [195].
One example, an improved RF resin, is a copolycondensate of amide-
substituted resorcinol and resorcinol-formaldehyde. [196].
(3) Improvement of rubber latex
Vinyl-pyridine, styrene, butadiene-terpolymer latex (VP) are effective in
adhesion and widely used for the latex in RFL. Similar to VP, other polar
groups are introduced to modify rubber latex. The copolymerization or
graft-polymerization of acrylonitrile, allyl amine, unsaturated ketone,
methacrylic acid and maleic acid is considered to be applicable for rubber
latex. [197-200].
(4) RFL recipe for special rubber
Adhesion becomes very difficult when special rubber compounds such
as CR, NBR, HR or EPDM are used instead of NR and SBR. The latex is
selected on the basis of the idea described in 2.2.4. One method is to use
CR latex for CR rubber and NBR latex for NBR rubber. [201, 177]. For
EPDM rubber, a proposed method uses a halogenized polyethylene dis
persion liquid and a mixed latex of BR, SBR and VP. [202, 203].
(5) Others
Other adhesives, surface active agents and reactive catalysts are added
to the RFL to improve adhesion. Several methods have long been in use.
One of them is a method in which carboxyl group substituted lignin is
added to the RFL. [204].
Also proposed is a method which uses an aminoplastoresin or reaction
product of epichlorophydrine and diamine under acidic state instead of
RF resin. [205-206].

2.2.5. Adhesive Treatment of Polyester


Adhesives for polyester may be classified by the nature of the functional
material:
(1) Isocyanates, blocked isocyanates
(2) Polyethylenereas
(3) Modified polyvinylchloride
(4) Polyepoxides
(5) Special phenol-formaldehyde systems
Classes (1), (4), and (5) are mainly used today. In this section, details of
each method will be explained.
TIRE CORD AND CORD TO RUBBER BONDING 97

Isocyanates and blocked isocyanates:


Isocyanate rubber cement is useful in polyester to rubber bonding as de
scribed before. The isocyanate group, -NCO, has a specific action for both
polyester and rubber. Once polyester is treated by isocyanate, which is
used as an organic solution, adhesion to rubber is facilitated remarkably
after RFL treatment. The process and experimental results with this
method are shown in what follows [207]:
Polyisocyanate ,
Polyester cord -» . - -» Drying -> RFL Treatment
in organic solvent

Resulting strength of adhesion


Isocyanate/RFL double dip system 36.3 Ib/in.
RFL single dip system 8.8 Ib/in.
Bond strength also depends on pH of the RFL. Satisfactory adhesion can
be achieved over a pH range of 9.8 to 10.7.
A further interesting feature is that cord treated with polyisocyanate can
be stored for prolonged periods under laboratory conditions, e.g. six
months, without fear of loss of adhesion. Patents were issued for these
methods [177, 178, 212]. They are simple and superior for adhesion, but
requirements of solvent recovery and ventilation, as well as toxicity, limit
their wide acceptance in the tire industry. Another drawback is its sensitiv
ity to water. Since isocyanates react easily with water and lose their activ
ity [213], they cannot be used in an aqueous adhesive system as is. Stabili
zation against water is accomplished by encapsulation [214], dimeri/.ation
[215-218], or blocking of isocyanate [215]. Utilization of blocked isocyan
ates is commoner. Phenol is widely accepted as a blocking agent. Phenol
blocked isocyanate is not decomposed further in water, and regenerates
free isocyanate when it is heated at 140 to 170°C. Some blocked isocyan
ates are sold on the market, for example, phenol blocked diphenyl-
methane diisocyanate is marketed by DuPont as Hylene MP.
The adhesive solution is prepared by mixing ball-milled Hylene MP, la
tex, and thickener. A mixture of Hylene MP and RFL is also a good adhe
sive for polyester [219-224].
Du Pont proposed the pretreating yarn process. In this case, melt ex
truded polyester fibers are cooled and treated with a special spinning prep
aration containing polyisocyanate, then drawn through a stream-jetted
chamber. [208] This type of pretreated polyester gives a good cord-to-rub
ber bonding with the same RFL treatment as nylon or rayon.
Polyethyleneureas:
Ethyleneurea is a derivative of an isocyanate, but does not act as an iso
cyanate generator [225]. Ethyleneurea is prepared through the addition re
action of ethyleneimine to isocyanate.
Originally ethyleneureas were used as an adhesive for wood [226], or
were incorporated into rubber cement as a bonding agent of cellulose fiber
or polyamide fiber to rubber [209, 210]. Application of ethyleneureas for
polyester-to-rubber bonding was developed by Japanese tire manufac
turers and fiber producers [222, 227-231]. They are available for both
single and double dip systems. A single dip system combined with RFL
was, however, developed purposefully.
As low-priced isocyanates are limited to tolylenediisocyanate and di
98 MECHANICS OF PNEUMATIC TIRES

phenylmethane diisocyanate, only aromatic ethyleneureas have practical


significance. They are insoluble in water and settle out easily.
Several ways have been developed to improve mechanical stability of
the resulting dip solution such as addition of thickener, introduction of
methylol groups into ethyleneureas [232], and so on [231]. With the thick
ened dip system, maturing conditions of the RF resin should be very care
fully controlled to avoid gelation of the dip solution [233-234].
Bond strength also depends on the heat treatment [34]. The ethyl-
eneurea-RFL single dip system had been developed in Japan, but it did
not find practical use in the tire industry because of poor dynamic per
formance at tire operating temperatures and inferior adhesion with high
vulcanization temperatures. As to the action of ethyleneurea with polyes
ter, Timmons speculated that it functioned the same as isocyanates, i.e.,
isocyanate should be regenerated from ethyleneurea [240]. But this does
not seem to be correct since ethyleneurea changes thermally to an isome-
rized product or a polymer, but never dissociates to an isocyanate and eth-
yleneimine [225]. Kigane's concept that the adhesion is dependent on re
action of polyester polymer chain ends and aziridine rings is fairly
acceptable [235].

Modified polyvinylchloride:
In 1958 a new aqueous double dip system was invented at CIL (Cana
dian Industries Limited) [236, 237]. The first dipping solution is coded
TR-5, and consists essentially of a poly(vinylchloride) latex and a reactive
polyamide. The TR-5 emulsion is nontoxic and its composition is as fol
lows:
Geon 151 (PVC 50% emulsion) 40.0
Versamide 125 5.0
Dioctylphthalate 12.0
Catanac SP 4.0
Triton X-100 0.6
Acetic acid 0.3
Isopropanol 2.8
Water 60.0
Versamide 125 is a low molecular weight polyamine resin containing
many free ammo groups.
High speed emulsifying equipment and specialized techniques are re
quired to prepare TR-5. Since recommended solid pickup of TR-5 is 0.5
percent concentration, the above concentrated TR-5 solution is diluted
7 : 1 with water just prior to use. The second dipping solution is RFL as
commonly applied to rayon and nylon.
Adhesion depends on the heat treatment, especially after the first dip
ping, see figure 2.59. Relatively low temperature heat treatment, 150° C,
after the second dipping gives satisfactory results [238]. This process was
used in some commercial polyester tire production in Canada, but it did
not find wide use. It is said that lack of wider acceptance was caused by
the necessity of using special techniques in preparing a stable emulsion
and by stiffness of the dipped cord [240],
Polyepoxides:
The original use of polyepoxide for rubber to polyester bonding is
found in patents secured by N. V. de Bataafsche Petroleum and Shell De
TIRE CORD AND CORD TO RUBBER BONDING 99

DWELL TIME AFTER FIRST DIP


(SECOND)

FIGURE 2.59 Effect of heat treatment condition after first dipping on adhesion.

velopment. Many other systems have been developed from these systems.
They may be classified into three groups shown in the following schemes,
see figures 2.60-2.62.
Group 1: The adhesive is composed of polyepoxide, hardener, and latex.
This class is exemplified by using the TRL-12 system of Du Pont [69].
First dip composition is the following mixture:
Glycidylether of glycerin 12.5 g
Lauryl sulfate 0.5 g
Water 125 cc
Gentac, 41% 50 g

GROUP 1
NV DE BATAAFSCHE PETROLEUM MAATSCHAPPIJ BRIT.H 78839 0998)

TOTO RAYON T. TAKEYiMAJ-d M ENDC

GRAFTED LA
AMINE HARDEN

OU PONT (MSCHEPARO) TRl 12 SYSTEM USP .130007 (1967) JARWP38-13JI6 0963)

PIRELLI SOCIET* PEB i BRIT. P. IP36.95I (1966)


OVA-80 CpPQ'.rMER

rOVO RAYON (T. TAKEYAMAs-d M ENOO) JAPAN Pit- 16.410 (1 96 7)

FIGURE 2.60 Flow sheet of dipping methods, Group 1 systems.


100 MECHANICS OF PNEUMATIC TIRES

GROUP 2
SMELL DEVELOPMENT (cw SCHRODER) U.SP 2.902.399 (19591

U.SP 3 J 07.96 6 (1967) JAPANp 42-1I.4S2 (!%7!

FIGURE 2.61 Flow sheet of dipping methods, Group 2 systems.

Second dip solution is prepared by mixing the following components:


Metaphenylenediamine 6.3 g
Water 138 cc
Gentac, 41% 195 g
Polyester cord is immersed in the first dip solution and then heat treated at
232° C for 1 .5 minutes. This treated cord is immersed in the second bath

CROUP 3

GOODY EAR (J » CAROINA) U.SP 3.247.043(1966) JAPANP 39-10514(1964)

DEEPING MILLIKEN RESEARCH COPPlWC PRUIT T aid *J SCHRQEOER) U.&P. 3.231*12(1966)

TOYO RAYON (E KATO T TAKEYAMA and M. ENOO ) JAPAN P. 42-3348(1967)


POLYEPOXIOE HEAT RESORCIN01- 1 HEAT
POLYESTER LATEX — TREATMENT HEAANETMVLENEl » TREATMENT
-TETRAMINEJ

TOYO RAYON (T TAKE YAMAand M ENOO) JAPAN P 42-16.40 1 0967)


POLYESTES OflAFTED LATEX HEAT RFL HEAT
> TREATMENT ' TREATMENT

FlOURE 2.62 Flow sheet of dipping methods. Group 3 systems.


TIRE CORD AND CORD TO RUBBER BONDING 101

and heat treated at 232° C for 1.5 minutes again. While the cord treated
only with the first dip shows relatively poor adhesion, 2 lb., the doubly
dipped cord has strong adhesion, 10 lb. This double dip system is superior
to the D-15 blocked isocyanate system for adhesion as shown in table
2.27. Drawbacks are lower adhesion at high temperature and difficulty
of handling in subsequent processing because of stickiness of the dipped
cord.
Group 2: All methods are double dip systems in which the first dipping
solution is a mixture of polyepoxide and hardener, and the second is RFL.
The D417 system of Du Pont containing isocyanate as hardener is a typi
cal example of this class. The composition of each dipping solution is tab
ulated as follows [261]:
First bath Second bath
Triton X-100 0.4 cc Resorcinol 73.7 g
Phenol blocked Formalin, 37% 40.0
diphenyl methane Vinyl pyridine latex,
diisocyanate 16.0g 41% 148
Water 400 cc Water 480
Diethylaminoethyl
methacrylate,
0.5% aq. sol 25 cc
Glycidylether of
glycerine 4.8 cc
Heat treatment after immersion in the first and second dip solutions was
218° C for 45 seconds.
A series of tests were conducted varying the proportion of the poly
epoxide and isocyanate from 0:100 to 100:0. Results are shown In fig
ure 2.63. For comparison ethylenediamine was used as a hardener. The re
sults are shown on the same graph.
It is well known that proper combination of the isocyanate and poly
epoxide increases the adhesion, and amine hardener is less effective under
these experimental conditions.
Although this system has many disadvantages such as requiring a
double dip, necessity of a time consuming ball-milling process, settling out
tendency of the dispersion, and lack of good dynamic performance at high
temperatures required for use in tires, many tires have been built using
this system, and it still provides a commercially available method [68].
Group 3: This class is also a double dip system. Polyepoxide is used in
the first dipping solution without hardener, and the second is RFL. An ex
ample of this class is Deering Milliken's process. Polyester cord is passed

TABLE 2.27 Comparison of dip systems

Dip TRL-12 Isocyanate RFL


Test method
Rubber NR SBR SBR SBR
Pull-through load, lb/0.25 in
Cold 28.8 22.0 19.3 16.0
Hot 10.8 9.5 7.4 6.0
Strip force, lb 4.2 7.9 6.3 1.5
102 MECHANICS OF PNEUMATIC TIRES

2 -

100 EPOXIOE
0 HARDNER

FIGURE 2.63 Effect ofpolyepoxide to hardener proportion on adhesion.

through the first dip solution. The recipe is as follows:


Vinyl pyridine latex, 41% 125 g
Eponite 100° 31.2
Petrowet 1.25
Water 312
" Eponite 100 is diglycidyl ether of glycerine (Shell).
Wet pickup is adjusted to give 3.3 percent solid pickup by passing
through a squeeze roll unit. The dipped cord is then passed into a heat
treatment oven at 232°C for 96 seconds. The resulting tacky cord is im
mersed in an RFL bath and heat treated again at 230° C for 120 seconds
with slight stretching. The second dipping solution is RFL which is pre
pared by maturing the following composition for 96 hours at room tem
perature:
Resorcinol 45 g
Formalin, 37% 22
Water 219
Isopropanol 5.8
Vinylpyridine latex, 41% 144
Sodium hydroxide, 10% Optional amount to
adjust pH to 9 to 9.3
The resulting cord is substantially free of stickiness and processes well.
Adhesion of this doubly treated cord to rubber is excellent. Table 2.28
gives the adhesion obtained at each step of this process. In this method the
adhesion depends upon the composition of the first dip solution and the
heat treatment. The proper range of concentration of polyepoxide gives a
high level of adhesion as shown in figure 2.64. Higher temperature treat
ment is preferable to give a high level of adhesion. This correlation is com
mon to most methods utilizing the polyepoxide. The following results
shown in figure 2.65 were obtained with IJeering Milliken's method.
A similar correlation is found with the polyepoxide-Versamide/RFL
double dip system of ICI (Belg. 662,282) and with the D417 system, i.e.,
polyepoxide-isocyanate/RFL double dip system of Du Pont [261].
TIRE CORD AND CORD TO RUBBER BONDING 103
TABLE 2.28 Comparison ofadhesion at each step ofDeering Mittiken's system
Pull-through
Step load, Ib
Untreated Oto4
Only first dipped 14 to 16
Only second dipped 10 to 12
Doubly dipped 27

The most popular epoxide for this purpose is glycidyl ether of glycerine
since it is easy to handle because of its water solubility. But it is necessary
to have regard for the life of the dipping solution because the epoxide
groups gradually undergo hydrolysis in an aqueous solution, and activity
drops by half in a week [34].
Compositions of RFL used for the second dip solution are either the
same for nylon and rayon or a special one for polyester. RF resin of lower
formaldehyde content is found in several patents for this purpose although
a lower formaldehyde RFL is inadequate for nylon and rayon. Examples
of special recipes are shown in table 2.30.
Polyepoxide pretreated polyester yarn:
Polyepoxide pretreated polyester yarn initially appeared in West Ger
many under names such as Trevira GPA (239), Diolen DSP (111), and
V75 (241).
Cords or fabrics of this type polyester give a good rubber to textile ad
hesion with the same RFL treatment as nylon or rayon. RFL treated Dio
len DSP cable cord (1000D3/3) gives a pull-through load of over 20kg/
cm., but the normal type only 10kg/cm. A comparison of two Diolen types
using two vinylpyridine latexes for the RFL component, and two adherent
rubber compounds is shown in table 2.3 1 .
In the manufacturing process for these modified polyesters undrawn
polyester fiber is treated with a spinning preparation which contains lubri
cant agents, polyepoxide, and amine hardener, then stretched and heat
treated at the same time.

CONCENTRATION OF POLYEPOXIDE IN
FIRST DIP SOLUTION (1.)

FIGURE 2.64 Effect of concentration ofpolyepoxide in first dipping solution on adhesion.


104 MECHANICS OF PNEUMATIC TIRES
1
FIRST SECOND
I

3 25 25
o

20
i
15

150 170 190 210 230 150 170 190 210 230
HEAT TREATMENT TEMPERATURE HEAT TREATMENT TEMPERATURE
PC) CO

FIGURE 2.65 Effect ofheal treatment temperature on adhesion of Peering MUUken's system.

It was thought that since existing dipping equipment could be employed


for these polyesters, the demand for them would grow if the adhesion was
satisfactory for tire cords. Spinning preparations are usually aqueous, so
the pretreating yarn process seems to be more adequate to the polyepoxide
system than the isocyante system.
However, down to the present, the pretreated type yarn has not been the
dominant type used for polyester tire cord.
CIL has developed a new polyester to rubber adhesive system [68, 240,
242-247]. It was called N3 and consisted of a reaction product of tri-
allylcyanurate, resorcinol, and formaldehyde of the following composi
tion.
Resorcinol 100 parts
Triallylcyanurate 24
Red lead catalyst 0.25
Formalin, 37% 28
Water 400
Ammonium hydroxide, 28% 28
N3 is available both for single dip and double dip processes. It is also ap
plicable to nylon.
Resorcinol to formaldehyde ratio of the RFL used with N3 affects the
adhesion of polyester to rubber, see table 2.32. For this adhesive system

TABLE 2.29 Effect ofheat treatment on polyester cord to rubber adhesion

Polyepoxide-Versamide/RFL system
Time, sec
Pull-through load,
kg/6.35 mm
First temp, °C 20 45 90
205° 7.12 7.53
234° 5.44 8.71 10.57
246° 7.85 9.53
260° 9.85 10.70 11.66
TIRE CORD AND CORD TO RUBBER BONDING 105

TABLE 2.30. Examples of RFL recipes for the second dip solution
A B C
Resorcinol/fonnaldehyde
molar ratio 1/0.66 1/0.68 1/0.73
Latex solid/RF solid
weight ratio 3.3/1 1.7/1 2.2/1
Resorcinol 48.4 10.3 10.0
Formalin, 37% 23.3 50.5 54
VP latex 40% 505 646
SBR latex 40% 445
6 Optional amount
Water 463 330 200
Source Goodyear Deering Milliken Du Pont
U.S. 3,247,043 U.S. 3,231,412 U.S. 3,307,966

high heat treatment temperature is necessary. Typical conditions are


drying at 104°C for 120 seconds and heat treatment at 254°C for 45 sec
onds. Addition of sodium carbonate or diethylenetriamine to the N3-RFL
mixture results in better adhesion after less severe heat treatment. Bonding
force also depends on adhesive pick-up and preferable pick-up is in a
range of 5 to 7 percent on the weight of cord.
ICI4 has developed a new single dip adhesive system. The additive is
called Pexul, formerly H7, and the dip solution is prepared by mixing Pe
xul with RFL of a special recipe. The composition of Pexul has not been
disclosed, but it can be surmised to be 2,6-Wj-(2,4′-dihydroxyphenylme-
thyl)-4-chlorophenol from the patent literature [249]. The working mecha
nism may be sorption of Pexul into polyester.
There are several single dip adhesives which could be attributed to sor
ption to polyester other than N3 and Pexul. They are furfural modified re-
sorcinol formaldehyde latex of AKU5 [248] alkylated resorcinol formal
dehyde latex [163] and I-RFL of Toray [38].
Recent improvement ofpolyester cord-to-rubber bonding:
Polyester cord is principally used for passenger car tires and the per
formance required for these tires is satisfactory with the adhesive systems
mentioned above.
In the case of heavy duty truck tires, further improvements of adhesion
under elevated temperature and dynamic conditions are required before
polyester can be widely accepted. The nylon cord/RFL system has been
used for truck and bus tires for many years. Therefore, it is believed that
Imperial Chemical Industries.
5 Algemene Kunspzijde Unie.

TABLE 2.3 1 Comparison of two Diolen types (111)

Pull-through load, kg/cm


Type of Diolen DSP Normal
Kind of VP latex A B A
Kind of rubber R D R D R D
Heat treated at 175 °C 15 22 23 25 8 10
Heat treated at 195 °C 16 22 26 29 7 9
106 MECHANICS OF PNEUMATIC TIRES

TABLE 2.32 Effect ofRFL composition on adhesion with the N3-RFL single dip system
A B C
Gentac latex, 41% 53.0 29.2 58.5
SBR latex 41% 73
Resorcinol 3.1 2.1 5.0
Formalin, 37% 4.9 3A 2.8
5.7 3.7
Water 33.3 54.3 33.7
Pull-through load, lb/0.25 in rubber compound
A IS 17 23
B 22 22 "27
C 12 19 19
" Rubber failure.

adhesion between RFL and rubber under severe tire operating conditions
is sufficiently stable. In the case of polyester, it becomes important to im
prove adhesion between the cord and adhesive at severe tire operating
conditions.
lyenger reported that the adhesion retained after heat aging cord-rub
ber composite samples at elevated temperatures was improved by eliminat
ing the sources of amines in the elastomers and compounding ingredients,
reducing the moisture content of the stocks, and introducing additives
such as 2-Chloro-4.6 diamino triazine to the elastomers [79, 80].
These methods are also used to reduce hydrolytic and am i noly tic degra
dation of physical properties of polyester cords. (See Section 2. 1 .4.). From
these facts, it appears that the degradation of polyester cord in rubber
stock is an important factor in the reduction of adhesion under severe heat
aging. Several other ideas for improving the adhesive bonding between
the polyester cord and RFL have been proposed. For example, the authors
have examined the three-step process, in which polyester cord was treated
initially with a polyisocyanate solution, subsequently with polyepoxide
solution, and finally with RFL dip. [266]. Superior adhesion retention in
heat aging tests was obtained by this method.
2.2.6. Adhesive Treatment of Miscellaneous Tire Cords
Steel Wire Cord: A special method has been used to bond rubber to
steel wire cord. A high level of adhesion can be obtained by vulcanizing
the rubber compound in contact with brass or zinc plated steel cord. Most
tire wire is plated at the wire mill, after drawing, by continuous elec
trolytic methods. Amount of brass and zinc on the wire is controlled in the
ranges of 4 to 8 g/kg., and 2 to 3 g/kg. respectively. It is mentioned that
the preferable brass composition is 70 percent copper and 30 percent zinc
[250]. Choice of the type of plating is decided by the nature of the rubber
compound. And it is said that adhesion is affected by ingredients of the
rubber compound [251-255].
Besides the plating method, addition of isocyanate or halogenated rub
ber adhesives to the rubber compound is also employed [135].
Glassfiber: Glass fiber cord is pretreated with an adhesive prior to treat
ing with RFL or rubber cement [86, 94]. The pretreatment is applied dur
ing the fiber forming process in combination with a variety of lubricants
which give improved fiber properties, especially abrasion resistance, to
glass fiber. Since unsaturated silancs such as vinylsilane are effective for
TIRE CORD AND CORD TO RUBBER BONDING 107

bonding glass to rubber, the pretreatment agent may contain such a com
pound [256-257].
Polypropylene: Polypropylene to rubber bonding is one of the most diffi
cult problems in this field. Several methods are patented. For example,
polyepoxide/RFL double dip was claimed to be effective [258]. Adhesion
of polypropylene for isocyanate rubber cement is improved by graft po
lymerization of vinylacetate, followed by hydrolysis [259]. It is mentioned
that polypropylene to ethylene propylene rubber bonding is attained sim
ply by vulcanization together if a peroxide curing system is used [260].
Vinylon: Vinylon is made bondable to rubber by the same treatments as
for nylon or rayon. Since Vinylon has a very hydrophilic nature, easy pen
etration of RFL adhesive into the fiber interior causes stiffening of the
dipped cord. Addition of water-repellent agents to RFL is used to reduce
this tendency [261]. Also pretreatment of Vinylon by softeners, a type of
surfactant, is effective for softening the dipped cord [262].
Wholly aromatic polyamide:
According to the technical information from Du Pont, "Kevlar" wholly
aromatic polyamide cord is treated with the polyepoxide/RFL double dip
system, the same as polyester. The heat treatment temperatures are also in
the same range as polyester cord [105, 1 16].
The following formulations are the typical adhesive systems, coded
IPD-22/D-5C.
IPD-22 (Predip solution)
Water 6.80
Sodium Hydroxide ( 1 0% Soln) 0.08
"Epon"812 0.16
2-Pirrolidone 0.80
"Aerosol" OT (5% Soln) 0. 16
D-5C (Top coat)
Mixture of D-5A RFL and 25% HAF block dispersion (9:1)
D-5A RFL
Water 55.60
Sodium Hydroxide ( 1 .67% soln) 4.475
Resorsinol 2.76
Formaldehyde (37%) 4.04
66.875
Water 15.28
"Gentac" Latex (41%) 6 1 .00
Ammorium Hydroxide 2.84
79.12
Total mixture 149.995
2.2.7. Evaluation of Adhesion
Adhesion has been evaluated by both tire tests and laboratory methods.
In both cases, measurements of the bonding strength and observation of
the failed state are important to judge the level of adhesion. And since ad
hesion falls gradually during tire operation, it is necessary to simulate a
change of adhesion with running of the tire. To meet the requirements,
108 MECHANICS OF PNEUMATIC TIRES

static measurements, dynamic measurements, and combinations of them


have been developed. However, there is no method of satisfactory evalua
tion on a laboratory scale.
Klein and co-workers [136] stated that stripping adhesion decreased
with mileage, and the highest adhesion loss occurred in the region of max
imum flexing (between 3.5 and 4 inches along the cord from the center of
the tire), figure 2.66.
Kenyon [263] mentioned that deterioration of adhesion of rayon tires
after running is greater than for nylon tires. Also, reduction of adhesion of
Ply 1 to 2 is greater than that of Ply 3 to 4 on four-ply tires.
Eccher [139] investigated the failed state by microscopic observation of
tire cross sections. He reported that damage was concentrated in the zone
beneath the tire shoulder, and consisted of rubber tears and separation be
tween cord and rubber.
Numerous methods have been proposed to estimate tire cord to rubber
adhesion. They are briefly explained here.
Static Tests
Pull-through type tests; H, T, and U-tests: Adhesion is represented by the
force required to pull an embedded cord through and out of a rubber
block in the H-test, T-test, and U-test. The various test names come from
the shapes of the test specimens, and are used only for convenience, see
figure 2.67.
In these methods, the force is affected by embedded length of cord, rate
of loading, and temperature [263, 264]. Measurement is frequently at 110°
to 130°C since temperature in a running tire may be this high.
I-test: The I-test is made by pulling both ends of an I-shaped test speci
men in which a sample cord is embedded as shown in figure 2.68. The
cord is represented by a dotted line. The force-deflection curve is re
corded. Adhesion is represented by the second peak force of figure 2.69.
Distortion of the I-test specimen during loading is shown in figure 2.68
[168].
Pop-test: The pop-test is made by compressing a cylindrical rubber test
specimen in which a sample cord is embedded in the middle as shown in
figure 2.70. The test specimen is deformed on application of load, from

TEST

TEST

0 2 4 6 B 10
DISTANCE, FROM CENTER OF CROWN (In.)

FIGURE 2.66 Deterioration of adhesion after fleet test and indoor wheel test.
TIRE CORD AND CORD TO RUBBER BONDING 109

CORD

MOLD

RUBBER

JIG

H TEST T TEST U TEST

FIGURE 2.67 Test specimens and testing features ofpull-through type test.

shape A to B. When failure of adhesions occurs, appearance of the speci


men changes to C or D. Adhesion is represented by the compression force
required to deform the test specimen to the C or D stage [265].
Cord stripping test: A form of stripping cord adhesion test is practiced in
Du Pont, the Single End Strip Adhesion (SESA) Test [21 1]. The force nec
essary to strip the cord from a rubber sheet is determined and is reported
as pounds force per single end of cord. The test arrangement is shown in
figure 2.71.
Fabric stripping test: Adhesion of tire fabric and belting duck is eval
uated by stripping a- two-ply specimen which is a sandwich structure of

100 ISO
ELONGATION

FIGURE 2.68 "/" test stress-strain curve compared with control sample with no cord [168].
110 MECHANICS OF PNEUMATIC TIRES

L71
ri

LJ LJ LJ
NO LOAD INITIATION
LJ LJ
PROPAWTION PROPAGATION PROPAGATION COMPLETE RECOVERED
OF FAILURE OF FAILURE BOTH ENDS TWIST FAILURE SAMPLE
ONE END ONE END

FIGURE 2.69 Distortion of "1" test specimen during load application [168]. Dotted line is the
cord.

rubber cover-test fabric-inner rubber-test fabric-rubber cover [267-268].


In evaluating tire fabric, high cord end counts are used to minimize rubber
bridge formation. The force is also affected by thickness of the inner rub
ber and rate of loading [269]. This method has been adopted to estimate
the level of adhesion in tire carcasses. Timmons [240] suggested that visual
inspection of strips frequently correlated with the appearance of cords in
tires after high speed tire testing. The testing arrangement is shown in fig
ure 2.72.
Dynamic: Adhesion gradually deteriorates with repeated deformation.
Dynamic evaluation of adhesion is made by counting cycles of deforma
tion to reach a limiting value. Testing equipment is commonly designed to
cut off when bonding force reaches a specific value. Type of deformation

ABC D
FIGURE 2.70 Deformation of Pop test specimen during load application [265].
TIRE CORD AND CORD TO RUBBER BONDING 111

FIGURE 2.71 Testing arrangement for Single End Strip Adhesion test.

FIGURE 2.72 Testing arrangement for two-ply strip test.


112 MECHANICS OF PNEUMATIC TIRES

STATIONARY
ANVIL

WE I 6 H T

FIGURE 2.73 Principle of Compression Adhesion Tester.

may be classified as: (1) compression, (2) shearing, (3) dynamic strip, and
(4) flexing.
Compression type: A test specimen is repeatedly compressed as illus
trated in figure 2.73. The test cord is under tension by means of a
weight. When the test specimen is compressed the rubber to cord interface
is subjected to shear force. Adhesion deteriorates with compression cycles,
and failure occurs when the bonding force becomes less than the weight
suspended on the cord. The Goodrich Flexometer has been conveniently
used for this purpose. Besides, the Goodrich Disk Fatigue Tester is avail
able, in which the specimen is subjected to both compression and exten
sion [272]. Compression-extension deformation may be used with the
equipment diagrammed in figure 2.74 [273]. A test cord is embedded in
the middle of a dumbbell shaped test specimen. In these tests, failure oc
curs just inside the rubber where there is severe heat degradation. There
fore, it is said that these methods are inadequate to examine adhesion
[187]. This method is available not only for cord but also for fabric [274-
276].

TEST SPECIMEN

FIGURE 2.74 Schematic drawing of Compression-Extension Type Tester and test specimen.
TIRE CORD AND CORD TO RUBBER BONDING 113

TEST SPECIMEN

SOCKET

ECCENTRIC
DRIVE

FIGURE 2.75 Schematic drawing of Dynamic Shear Adhesion Tester.

Shearing type: Continuous vibration is applied to a rubber specimen as


illustrated in figure 2.75. This is an example of this type testing equipment,
the Dynamic Shear Adhesion (DSA) Tester. Adhesion is represented by
cycles to failure of the cord-to-rubber bond and this is automatically re
corded [277]. ,
Dynamic strip type: In the case of tire fabrics or belting ducks, a two-ply
strip testpiece is subjected to continuous vertical vibration under a certain
weight. Adhesion is expressed by cycles required to strip a unit length.
This type of equipment has been on the market and called a Scott Tester,
see diagram in figure 2.76.

FIGURE 2.76 Schematic drawing of Dynamic Strip Adhesion Tester (Scott Tester).
114 MECHANICS OF PNEUMATIC TIRES

ECCENTRIC DRIVE

UPPER ANVIL
i*

LOWER ANVIL

.TEST SPECIMEN

FIGURE 2.77 Principle of Dynamic flex Adhesion Tester.

Flexing type: Dipped cord to be evaluated is embedded in a cylindrical


rubber specimen. Flexing is by deformation of the specimen between a
stationary anvil and a plunger, see figure 2.77 [278]. Compression defor
mation can be given to the specimen at the same time by adjustment of the
position of the stationary anvil. An alternative flex adhesion fatigue testing
equipment was proposed. The test strap has a band of cords passing longi
tudinally through the middle. Affixed at one end to a stationary mounting,
the strap passes horizontally from there to a pair of little rollers which are
mounted in a reciprocating carriage. The strap follows a sigmoidal path
around the rollers, and continues horizontally to its free end on which
hangs a weight. Figure 2.78 gives a schematic view of the equipment. Ad
hesion is evaluated by a comparison of pull-through load before and after
flexing [279].
In evaluating fabrics, the test specimen is built as a two-ply sandwich
structure. Compressive fatigue is produced by flexing it over a spindle.

CARRIAGE

FIGURE 2.78 Schematic side view of Roller Flex Adhesion Tester.


TIRE CORD AND CORD TO RUBBER BONDING 115

SPINDLE

WEIGHT

TEST .
STRAP

FIXED
END

ECCENTRIC
DRIVE —

FIGURE 2.79 Schematic drawing of Belt Flex Type Dynamic Adhesion Tester.

The principle is illustrated in figure 2.79 [147, 280]. Adhesion is estimated


by either comparison of strip adhesion before and after flexing or cycles
for ply separation [282].

References
1] Moncrief, R. W., Man-Made Fibers, p. 234 (John Wiley & Sons, New York, 1963).
2] Matsudaira, N., Kaut. Gummi Kunstst. 19, 156 (1966).
3] Matsudaira, N., J. Soc. Rubber Ind. Japan 39, 289 (1966).
4] Kovac, E. J. Tire Technology. The Goodyear Tire & Rubber Co. 1973.
5] Ewald, G. W., Rubber Age 98(3), 57 (1966).
6 Chem. Eng. News 43(37), 41 (1965).
7] Chem. Week 97(24) (1965).
8 Ebert, A., Kaut. Gummi Kunstst. 18, 372 (1965); Rubber World 152(6), 98 (1965).
9] Curley, J. B., Rubber World 156(6), 58 (1967).
[10 Cox, N. U U.S. Patent 2,536,014 (1950).
[11 Röder. H. L., J. Textile Inst. 46, p. 84 (1955).
[12 Warzee, M., and Quintelier, G., J. Textile Inst. 46, p. 123 (1955).
[13 Howsman, J. A., and Sission, W. A., in Cellulose and Cellulose Derivatives, Part I, Ed.
E. Ott, p. 254 (Interscience Publishers, New York, 1964).
[14] Maeda. H., J. Soc. Textile Cellulose Ind. Japan (Sen-i Gakkaishi) 12, 6 (1956); Ko-
bunshi (High Polymer, Japan) 4, 413 (1955); 7, 261 (1958).
[15] Sobue, H., and Fukuhara, S., J. Chem. Soc. Japan, Ind. Chem. Section 59, 990 (1956);
60,323(1957).
[16] Mitsuishi, Y., and Maeda, H., J. Soc. Textile Cellulose Ind. Japan (Sen-i Gakkaishi)
18, 1049 (1962).
[17a Recent Development in Tire Yams, American Enka Corp., 1963.
[17b Fukuhara, S., Kobunshi (High Polymer, Japan) 17, 7 (1968).
[17c] Fukahara, S., J. Soc. Rubber Ind. Japan 40, 39 (1967).
18] Mukoyama, E., J. Textile Machinery Soc. Japan, Proc. 20, 706 (1967).
19] Hannell, J. W., Akron Rubber Group Symp., Akron, Ohio (Jan. 23, 1970).
20] Chemstrand, Fr. 1,323,811 (1968); Belg. 615,991 (1962).
21] Du Pont, Brit. 811,349 (1959); Brit. 889,144 (1962).
22] Monsanto, Brit. 955,903 (1964).
1 16 MECHANICS OF PNEUMATIC TIRES

[23a] Toyo Rayon, Japan 42-25, 502 (1967).


[23b] Toyo Rayon, Japan, 42-27, 572 (1967).
[24] Hattori, H., Kobunshi (High Polymer, Japan) 11, 481 (1962).
[25] Hattori, H., Kobunshi (High Polymer, Japan) 12, 615 (1963).
[26] d'Albignac, J., J. Soc. Dyers Colour. 82, 389 (1966).
[27] Howe, D. E., Du Pont Tire Yarn Tech. Rev., p. 5 (1957); Patterson, R. G., McCrea, H.
H., and Howe, D. E., Rubber World 138, 409 (1958).
[28] Howe, D. E., Du Pont Tire Yam Tech. Rev., p. 7 (1960).
[29] Bolmeyer, J. W., Du Pont Tire Yarn Tech. Rev., p. 5 (1958).
[30 Litzler, C. A., Southern Rubber Group, Dallas, Texas (Feb. 22, 1963); Rubber World
148(2), 68(1963).
31] Toyo Rayon, Toray Tire Cord News, No. 3 (1966).
32] Toyo Rayon, Toray Tire Cord News, No. 4 (1967).
33] Firestone, Brit. 809,916 (1959); Brit. 822,809 (1959); U.S. 2,955,344 (1960); U.S.
2,955,345 (1960).
[34] Toyo Rayon, unpublished data.
35] Goy, R. S., and Möring, P. L. E., Trans. Inst. Rubber Ind. 40, 176 (1964); Goy, R. S.,
E. S., Textile Inst. Ind. 5, 220 (1967); Goy, R. S., and Möring, P. L. E., Rubber
World 152(3), 67 (1965).
36] Pieper, E., Kaut. Gummi Kunstst. 16, 18 (1963).
37] Reegen, S. L., and Sabo, J., J. Appl. Polymer Sci. 2, 337 (1959).
38] Toyo Rayon, Toray Tire Cord News, No. 5 (1968).
39 Toray Industries Inc. Toray Tire Cord News No. 6 (1969).
40 Teheyama, T. and Higiri M. Reprint of 6th Kwansai Seni Seminar OSAKA (1973).
41 Dismore, P. F., and Station, W. O., J. Polymer Sci. C13, 133 (1966).
42 Kujimoto, K ., and Fukuda, K , Annual Meeting of the Society of Rubber Industry, Ja-
pan, Fukuoka, Japan, May 16, 1968.
[43] Todohi, M. and Kawagushi T., Kobunski Ronbunshu. Eng. Ed., 3, 1217 (1974), 3 1616
(1974), 4 139 (1975).
[44] Hechert, D. H., Du Pont Tire Yam Tech. Rev., p. 14 (1959).
45] Mihell, W. G., Du Pont Tire Yarn Tech. Rev., p. 37 (1958).
[46] Howard, W. H., and Williams, M. L., Rubber Chem. Tech. 40, 1139 (1967).
[47] Tipetts, E. A., Textile Research Institute, 36th Annual Meeting, New York, April 1,
1966.
[48] Rye, G. W., and Martin, J. E., Rubber World 149(1), 75 (1963).
[49] Papero, P. V., Wincklhofer, R. C., and Oswald, H. J., Rubber Chem. Tech. 38, 999
(1965).
[50] Claxton, W. E., Forster, M. J., Robertson, J. J., and Thurman, G. R., Textile Res. J.
36, 903 (1966).
51] Du Pont, Brit. 918,637 (1963); Japan 40-17,083 (1965).
52] Allied Chemical, Belg. 675,292 (1965); South Africa 918,637 (1963).
53 Monsanto, Belg. 665,541 (1965).
54 U.S. Rubber, Brit. 893,386 (1965).
55! Swanson, J. M., Du Pont Tire Yarn Tech. Rev., p. 37 (1958).
56 Brownlee, J. L., and Perry, E., Rubber Chem. Tech. 40, 1 147 (1967). . ,.
5f Mukoyama, E., and Takeyama, T., Kobunshi (High Polymer, Japan) 17, 14 (1968).
58 Cipriani, C., Papero, P. V., and Morre, M. S., J. Appl. Polymer Sci. 10, 601 (1966).
59' Chemiefasern 16,421 (1966).
60 Chem. Week 98(10), 35 (1966).
61 Du Pont, U.S. 3,220,456 (1965); Japan, p. 41-7889 (1966).
62
631 Du Pont U.S. 3,258,049 (1965).
Chem. Eng. News 43(37), 41 (1965).
64J Firestone, U.S. 3,378,602 (1968).
65] Kovac, F. J., Rye, G. W., and Dague, M. F., Ind. Eng. Chem. Prod. Res. Develop. 2,
279 (1963).
66] McCune, L. K., Textile Res. J. 32, 762 (1962).
67] Sprague, B. S., and Singleton, R. W., Textile Res. J. 35, 999 (1965).
[68] Aitken, R. G., Griffith, R. L., Little, J. S., and McLellan, J. W., Rubber World 151(5),
58 (1965).
69] Kovac, F. J., and Kersker, T. M., Textile Res. J. 34, 69 (1964).
70] Kovac, F. J., and McMillen, C. R., Rubber World 152(5), 83 (1965).
71] Ishixaki, S., J. Textile Machinery Soc. Japan 20, 696 (1967).
Craig, R. A., Du Pont Tire Yarn Tech. Rev., p. 21 (1960).
Du Pont, Japan 41-7892 (1966).
For example, Manabe, T., J. Textile Machinery Soc. Japan 8, 685 (1955).
Du Pont, Japan 37-5821 (1962).
TIRE CORD AND CORD TO RUBBER BONDING 117

76 Ravens, D. A. S., and Ward, I. M., Trans. Faraday Soc. 57, 150 (1961).
77 Ravens, D. A. S., Polymer 1, 375 (1960).
78 Little, J. S., Akron Rubber Group Meeting, Akron, Ohio (Jan. 23, 1970).
79 lyenger, Y., J. Appl. Polymer Sci. 15 267 (1971).
[80 lyenger, Y. and Ryder D. F., Rubber Chem. Tech. 46 422 (1973).
[81] Bukowski, R. A., J. Elastomers Plastics, 6 223 (1974).
ICI. US. 3692711 (1972).
Toray Industries, Toray Tire Cord News. No. 7 (1972).
84) Ryckebosch, F., Trans. I.R.I. 38, 79 (1962).
85] Beckaert Steel Cord Technical Information.
Litzler, C. A., Rubber World 141, 229 (1959).
Josifek, C. W., Kaut. Gummi 19(3), 160 (1966).
Miles, D. C., Rubber Plastic Age 45, 521 (1964).
DeBruyne, E. A., and Vanhoutte, J. S., J. Appl. Polymer Sci. 17 2033 (1972).
90] Zouck, J., Wire Journal, Apr. 1972, 41.
91] Del Gatto, J. V., Rubber World, Nov. 1970, 47.
[92] Monsanto, U. S., 3216076 (1965), U.S. 3163158 (1971), Brit. 1153577 (1969), Brit.
1210920 (1970).
P3 Marzocchi, A., and Gagnon, R. K , Rubber World 156 55 (1967)
[94 Marzocchi, A., and Leary, D. E., Am. Dyestuff Reporter 54, 870 (1965).
[95 Marzocchi, A., and Rooney, M. J., J. Polymer Sci. C19, 227 (1967).
[96 Wilson, J. B . Rubber Plastics Age 45, 525 (1964).
[97 Welch, F. G., Rubber World 156(5), 60 (1967).
[98 Ralph, F. W., Rubber Age 99(9), 59 (1967).
[99 Rubber Age 100(2), 100 (1968).
100 Rubber World 154(1), 34, 37, (1966).
101] Rubber Age 98(4), 86 (1966).
102] Kerrivan, A. R., J. Appl. Polymer Sci. 17 2027 (1973).
\ 103 PPG Ind., Tech. Information. Bulletin F-200.
104] Lachut, F., Rubber World. May 1972, 29.
105] Wilfong, R. E., and Zimmerman, Rubber World, July 1972, 40, J. Appl. Polymer Sci.,
17 2039 (1973).
106] Little, J. S., Akron Rubber Group Symp., Akron, Ohio (Jan. 23, 1970).
107] Kawai, K., Rubber Age, Jan. 1971, 71.
108] Tsuboi, K., 7th Kwansai Seni Seminar, Osaka, (Mar. 28, 1974).
109] Wolf, R. F., Rubber Age, March 1970, 83.
110] Alliger, C., SAE Journal, May 1973, 56.
Ill] Schroeder, W. A., and Prettyman, I. B., Rubber Age 99(1), 72 (1967).
112] Daimler, B. II., Kaut. Gummi Kunstst. 18, 15 (1965).
1 13] Promislow, A. L., Steele, R. and Whitworth, P. D., J. Appl. Polymer Sci. 17. (1973).
1 14] Draves, C. /.. Jr., Lee, Z. S., Skolnik, L., A.C.S. Rubber Division Meeting, Chicago,
Illinois (Oct. 20, 1970).
[1 15] Rodenkirch, B. L., 73rd National Meeting of the American Institute of Chemical Engi
neers. Minneapolis, Minnesota (Aug. 28, 1972).
116] Litzler, C. A., Rubber Age, Feb. 1973, 27.
117] H* all, I. H., J. Appl. Polymer Sci. 8, 237 (1964).
118] Lothrop, E. W., Appl. Polymer, Symp. No. 1, 111 (1965).
119] Lothrop, E. W., Appl. Polymer Symp. No. 5, 53 (1967).
120] O'Neil, K. B., Dague, M. F., and Kimmel, J. E., Appl. Polymer Symp. No. 5, 41
(1967).
[121] Lyons, W. J., Textile Res. J. 32, 448 (1962); 32, 553 (1962).
[122] Prevorsk, D. C., and Lyons, W. J., Textile Res. J. 85, 1 10 (1965); 85, 217 (1965); Lyons,
W. J., and Prevorsk, D. C., Textile Res. J. 35, 1 109 (1965); Lyons, W. J., and Rib-
nick, A. S., Textile Res. J. 37, 1014 (1967).
[123] Wilson, M. W., Tappi 43, 129 (1960).
[124] Lessig, E. T., U.S. 2,240,505 (1941); Busse, W. F., Lessig, E. T., Loughborough, D. L.,
and Larrick, L., J. Appl. Phys.
'hys. 13, 715 (1942).
125 Rosevear, W. E., and Waller,, R. C., Textile Res. J. 19, 633 (1949).
126' Lyons, W. J., Textile Res. J. 32, 750 (1962).
127 Lessig, E. T., U.S. 2,291,086 (1942).
128 Budd, C. B., and Larrick, L., U.S. 2,488,761 (1949); Budd, C. B., Textile Res. J. 21, 174
(1951).
[129] Noshi, H., Murata, A., and Hirata, Y., J. Textile Machinery Soc. Japan 15,46 (1962).
[130] (a) Fujino, K., Noshi, H., and Amau, M., J. Textile Machinery Soc. Japan 6, 626
(1953); (b) Fujino, K.., Noshi, H., and Yuse, M., ibid. 8, 254 (1955); (c) ibid. 8, 254
(1955); (d) Fujino, K., Noshi, H., and Shiomi, A., ibid. 9, 164 (1956); (e) ibid. 9, 334
118 MECHANICS OF PNEUMATIC TIRES

(1956); (0 ibid. 9, 378 (1956); (g) ibid. 10, 689 (1957); (h) Fujino, K., Noshi, H., and
Matsubayashi, F., ibid. 11, 100 (1958).
[131] Mitome, N., Manabe, T., and Tojima, T., J. Textile Machinery Soc. Japan 9, 97 (1956);
9, 248 (1956).
[132] Frank, F., and Singleton, R. W., Textile Res. J. 34, 11 (1964).
[133] Prevorsk, D. C, and Lyons, W. J., Textile Res. J. 34, 881 (1964); 34, 1040 (1964).
[134] Noshi, H., Fujimoto, H., and Yoshida, T., J. Textile Machinery Soc. Japan 14, 256
(1961).
[135] Williams, K.. R., Hannell, J. W., and Swanson, J. M., Ind. Eng. Chem. 45, 7% (1953).
[136] Klein, W. G., Platt, M. M., and Hamburger, J. J., Textile Res. J. 32, 393 (1962); Tappi
43, 657 (1960).
[137] Patterson, R. G., and Anderson, R. K., Rubber Chem. Tech. 38, 832 (1965).
[138 Patterson, R. G., Rubber Chem. Tech. 39, 1382 (1966).
[139 Eccher, S., Kaut. Gummi Kunstst. 19, 299 (1966); 19, 372 (1966).
[140 Mieck, K.-P., and Lauchner, W., Faserforsch. Textiltechn. 16, 16 (1965).
[141 Miyamoto, S., Washimi, Y., and Fujimoto, N., J. Soc. Rubber Ind. Japan 38, 48
(1965).
[142] Miyamoto, S., Washimi, Y., Nagai, H., and Fujimoto, N., J. Soc. Rubber Ind. Japan
39, 53 (1966).
[143] Wood, J. O., and Redmond, G. B., J. Textile Inst. 56, T191 (1965).
[144] Dillon, J. H., Fatigue phenomena in high polymers, in Advances in Colloid Science,
Vol. III, Ed. Mark, H., p. 219 (Interscience Publishers, New York, 1950).
[145] Mallory, G. D., U.S., 241,524 (1946).
[146] AS I'M Standards, Part 24, Textile Materials, Method D885-64, Appendix, In-rubber
fatigue determination of textile tire cords, p. 175 (American Society for Testing and
Materials, Philadelphia, 1967).
Kenyon, D., Proc. I.R.I. 11, 67 (1964).
Wilson, M. W., Textile Res. J. 21, 47 (1954).
Noshi, H., and Shiroguchi, T., J. Textile Machinery Soc. Japan 16, 828 (1963).
Mieck, K.-P., Faserforsch. Textiltechn. 10, 578 (1959).
Kainradl, P., and Handler, F., Faserforsch. Textiltechn. 11, 408 (1960).
Kern, W., Kaut. Gummi 8, WT195; WT233 (1955).
Mieck, K.-P., Plaste Kaut. 7, 344 (1960).
Kemmnitz, G., and Espanion, G., Chemiefasern 16(3), 182 (1966).
Kemmnitz, G., J. Appl. Polymer Sci. 6, 373 (1962).
Fujimoto, S., J. Textile Machinery Soc. Japan 18, P57 (1965).
Crocker, G. J., Rubber Chem. Tech., 42, 30 (1969)
Hultzsch, K., Kunstist. 37, 43 (1947).
Greth, A., Angew. Chem. 51, 719 (1938).

162 Mather, Br. Polymer J. 3 (3) (1971).


163 lyenger, Y., and Erickson E., J. Appl. Polymer Sci. 11 2311 (1967).
164 Dietrick, M. I., Rubber World 136 847 (1957).
165] Iyenger, Y., Rubber World 148 (6) 39 (1963).
Fukuzawa, High Polymers, Japan 19453 (1970).
General Tire brochure, Gentac.
Doyle, G. M., Trans. I.R.I. 36, 177 (1960).
Dubrisay, R., and Papault, R., Compt. Rend. 215, 348 (1942).
Le Bras, J., and Piccini, I., Ind. Eng. Chem. 43, 381 (1951).
Uzina, R. V., et al. Sov. Rubber Tech. 20(7), 18 (1961).
Levitin, I. A., et al, Sov. Rubber Tech. 21(1), 41 (1962).
174 van Giles, G. E., Polymer Preprints 8(1), 508 (1967).
175 Du Pont, Brit. 1,106,920 (1968).
176 Esso, Brit. 803,182 (1958); Brit. 823,282 (1959); Brit. 861,814 (1961).
177 Burlington, U.S. 3,240,659 (1966).
178 Burlington, U.S. 3,240,649 (1966).
179 Stamicarbon, Belg. 680,492 (1966).
180 Dunlop Rubber, U.S. 3,268,386 (1966).
1811 Du Pont, U.S. 3,276,948 (1966).
182] Farbwerke Hoechst, Japan 42-17642 (1967).
183] Montecatini, Japan 42-23632 (1967); Brit. 1,071,528 (1967).
184] Reeves, L. W., Rubber World 132, 764 (1955).
185] Reeves, E. V., Adhesive Age 6(3), 28 (1963).
186] liannel, J. W., Du Pont Tire Yam Tech. Rev., p. 27 (1958).
TIRE CORD AND CORD TO RUBBER BONDING 119

187 James, D. I., Wake, W. C, Trans. I.R.I. 39, 103 (1963).


188 FMC Corporation, Japan 42-6668 (1967).
189 Murphy, R. T., Baker, L. M., and Reinhardt, Jr., R., Ind. Eng. Chem. 40, 2292 (1948).
190 Gillraan, H., and Thoman, R., Ind. Eng. Chem. 40, 1237 (1948).
[191] Iyenger, Y., J. Appl. Polymer Sci. 13 353 (1969).
[192 Toray, Japan 45-260 (1970).
[193 Asahi Chemical, Japan 46-42612 (1971).
[194 Asahi Chemical, Japan 46-41089 (1971).
[195 Yokohama Rubber, Japan 45-20599 (1970).
[196 Ashland Oil & Refine, US 3410818 (1968).
[197 Dunlop, Brit, 1112557 (1968).
[198] Dunlop, US 3396065 (1968).
199] Dunlop, Japan 45-26439 (1970).
200] Toyo Rubber, Japan 458341 (1970).
201] Burlington, US 3240660 (1966).
202] Asahi Chemicalm Japan 45-23142 (1971).
203] Uniroyal, Ger. Offen, 2039692 (1971).
204] USSR 275372 (1970).
205] Am. Cyanamid, US 3361617 (1968).
[206] USSR 276395 (1970).
207] Meyrick, T. J., and Watts, J. T., Proc. I.R.I. 13 52 (1966).
208] DuPont, Japan 43-21507 (1968).
209] Hodogaya Kagaku, Japan 32-8192 (1957).
210] Kaneko, T., Japan 30-1895 (1955); Japan 31-1692 (1956).
211] DuPont, U.S. 3307966 (1967).
212] ICI, Japan 35-13125 (1960); Brit. 816,640 (1959).
213] For example, Shkapenko, G., Gmitter, G., Gruber, E., Ind. Eng. Chem. 52, 605 (1960).
214] Farbwerke Hoechst, Japan 41-12109 (1966).
215] Bayer, O., Angew. Chem. A59(9), 257 (1947).
216] Tanabe, T., et al., Kogyo Kagaku Zasshi (J. Chem. Soc . Japan, Ind. Chem. Sec.) 66,
821 (1963).
[217 Farbwerke Hoechst, Belg. 656,656 (1965); Belg. 672,260 (1966).
[218 Yokohama Rubber, Japan 42-784 (1967).
[219 Thompson, W. L., and Parke, L. W., Rubber World 138, 588 (1958).
220 Du Pont, U.S. 2,994,671 (1961).
221 Du Pont, Instruction Manual for D-15, 1957.
222] Teijin, Japan 38-13236 (1963).
223] Goodyear Tire and Rubber, Japan 40-21429 (1965); U.S. 3,268,467 (1968); Japan 40-
10723 (1965); U.S. 3,226,276 (1966).
[224] Toyo Rayon, Japan 38-20961 (1963).
[225] Iwakura, Y., Nabeya, A., and Nishiguchi, T., Yuki Gosei Kagaku Kyokai Shi (J. Soc.
Organic Synthetic Chem. Japan) 26, 101 (1968).
[226] Iwakura, Y., in Setchaku, High Polymer Series 1, p. 185 (High Polymer Soc. Japan ed .
Maruzen, 1965).
Toyo Rubber, Japan 42-2896 (1967).
Toyo Rayon, Japan 42-2275 (1967); Japan 40-26548 (1965).
Teijin, Japan 37-3123 (1962); Japan 37-3124 (1962).
Toyo Spinning, Japan 42-543 (1967).
Takeda Pharmacy, Japan 42-8221 (1967); Japan 42-541 (1967); Japan 42-860 (1967).
Teijin, Japan 39-10513 (1964); Japan 40-28937 (1965); Japan 41-5260 (1966); Brit.
953,415 (1964), U.S. 3,235,333 (1967); I.C.I. Belg. 674,529 (1966).
Teijin, Japan 40-922 (1967).
Mitsuboshi Belt, Japan 40-16135 (1965).
Kigane, K., Japan 38-20956 (1963).
Little, J. S., Can. Textile J. 78(19), 57 (1961).

238 Chapman, R., Rubber Plastic Age 47, 769 (1966).


239 The Daily News Record, Aug. 17, 1965.
240 Timmons, W. D., Adhesive Age 10 (10) 27 (1967).
241 Daimler, B., Kaut. Gurnmi Kunstst. 20 (3) 159 (1967).
242 Vereingte Glanzstoff-Fabriken, Brit. 1,025,310 (1966); U.S. 3,297,467 (1967); Brit.
1,035,299 (1916); Japan 42-19921 (1967); Brit. 1,025,309 (1966); U.S. 3,297,468
(1967); Brit. 1,035,220 (1966); Japan 42-19922 (1967); Brit. 1,012,935 (1966); Brit.
1,026,221 (1966).
[243] Canadian Industries Ltd., U.S. 3,318,750 (1967).
[24 ] ICI, Be g. 674,528 ( 966).

[245] Smolin, E., and Rapoport, L., The Chemistry of Heterocyclic Compounds, S-Triazines
and Derivatives, p. 400 (Interscience Publishers, New York, 1955).
[246] Clampitt, B. H., German, D. E., and Galli, J. R., J. Polymer Sci. 27, 515 (1958).
[247] Saunders, J. H., and Frish, K. C., Polyurethanes Chemistry and Technology, Part 1, p.
116 (Interscience Publishers, New York, 1962).
[248] Algemene Kunstzijide Uni, Fr. 1,481,392 (1967); Belg. 682,153 (1966).
[249] ICI, Belg. 688,424 (1967); Fr. 1,496,951 (1967); Neth. 6,614,669 (1967); Belg. 688,425
(1967).
250] Wake, W. C., Reference [175], p. 413.
251] For example, Hicks, E., and Lyon, F., Rubber Chem. Tech. 40, 1607 (1967).
252 Adler, O. E., Adhesive Age 12(1), 30 (1969).
253 Hicks, A. E., and Lyon, E., Adhesive Age 12(5), 21 (1969).
254 Sexsmith, F. H., Adhesive Age 13(5), 21 (1970).
255 Sexsmith, F. H., Adhesive Age 13(6), 31 (1970).
256 Vanderbilt, B. M., and Clayton, R. E., Rubber Chem. Tech. 38, 379 (1965).
257 Esso, U.S. 3,307,967 (1967).
258 Toyo Rayon, Japan 42-3947 (1967).
259 Toyo Rayon, Japan 42-542 (1967).
260 Montecatini, Japan 42-26830 (1967).
261 Kurashiki Rayon, Japan 36-7742 (1961); Japan 37-12119 (1962).
262] Mitsuboshi Belt, Japan 36-7742 (1961).
263] Kenyon, D., Trans. I.R.I. 38, 165 (1962).
264] Wood, J. O., Rubber Chem. Tech. 40, 1014 (1967).
265] Lessig, E. T., and Compton, J., Rubber Chem. Tech. 19, 223 (1946).
266] Toray Ind. Japan Kokai 51-70394 (1976).
267] JIS K-6301-1962, Physical Testing Methods for Vulcanized Rubber.
268] ISO Recommendation R36, Rubber Chem. Tech. 31(1), xxi (1958).
269] Barroff, E. M., Elliot, R., and Wake, W. C., Rubber Chem. Tech. 25, 391 (1952).
270] Uzina, R. V., Gromova, L. S., Rubber Chem. Tech. 32, 898 (1959).
271] Gardner, E., and Williams, P., Trans. I.R.I. 24, 284 (1949).
272] Kenyon, D., Trans. I.R.I. 40, 67 (1964).
273] Khromov, M. K., Reznikovskii, M. M., and Lazareva, N. K., Sov. Rubber Tech. 21(6),
25 (1962).
[274] Levitin, I. A., Korablev, Yu. G., Komev, A. E., and Babitskii, B. L., Rubber Chem.
Tech. 32,889(1959).
[275] Kawai, K., Akron Rubber Group Symp. Akron, Ohio (Jan. 23, 1970).
[276] Tsydik, M. A., Lukomskaya, A. I., and Slonimskii, G. L., Rubber Chem. Tech. 33 42
(1960).
277] Miller, A. L., and Robinson, S. B., Rubber World 137 397 (1957).
278] Pittman, G. A., and Thornly, E. R., Trans. I.R.I. 25, 116 (1949).
279] Lyons, W. J., Anal. Chem. 23, 1255 (1951).
280] ASTM Standards, Part 28, Rubber, Carbon Black; Gaskets, Method D430-59, Dy
namic testing for ply separation and cracking of rubber products, p. 233 (American
Society for Testing and Materials, Philadelphia, 1968).

Blocked isocyanate : An isocyanate derivative that decomposes to an isocyanate


upon heating. Phenol-blocked aromatic di- or tri- isocyanates
are used as adhesives for rubber-to-polyester bonding.
Cord : Two or three plied yarn twisted together in one operation (also
called a cable yarn).
Denier : A number to indicate linear density of yarn or cord. The
weight in grams of 9000 meters of the yarn or cord.
Epoxide : A compound containing expoxide radical di-or tri- expoxides
are used as adhesives for rubber-to-polyester bonding.
^CH
Ethylene urea : A compound containing -NHCON\^i 2 radical
CH2
Filament : A fiber of infinite length.
Isocyanate : A compound containing the isocyanate radical -NCO. Aro
matic di-or tri- isocyanates are used as adhesives for rubber-
to-polyester bonding.
Nylon : Generic name for a polyamide family characterized by the
presence of the amide group -CON H The most important
are nylon 6 and nylon 66.
TIRE CORD AND CORD TO RUBBER BONDING 121

Polyester fiber : Generic name for a synthetic fiber in which the fiber forming
substance is any long chain polymer composed of a dihydric
alcohol and terephthalic acid. The most important is polyeth
ylene terephlhalate.
Rayon : Generic name for a manufactured fiber composed of regener
ated celloluse.
RFL : A mixture of rubber latex and resoreinol-formaldehyde resin.
Basic tire cord adhesive.
Tenacity : The maximum force developed in a tensile test taken to rup
ture, expressed in grams per denier, g.p.d. or G/D.
Wholly aromatic polyamide: Generic name for a polyamide family characterized by the fol
lowing chemical formula:
[NH - R - NHCO - R' - CO-]0
or
[CO - R - NH]n
R and R′ : Aromatic radicals.
Chapter 3
CORD REINFORCED RUBBER
J. D. Walter1

3.1.1. Introduction and Historical Perspective 124


3.1.2. Constituent Material Properties 126
Rubber Elastic Constants 126
Cord Elastic Constants 129
3.1.3. Cord-Rubber Composite Properties 130
One-Ply Systems: Engineering elastic constants 130
One-Ply Systems: Compliance and Stiffness Constants 141
Multi-Ply Systems: Classical Lamination Theory 146
Multi-Ply Systems: Interlaminar Deformations 157
3.1.4. Applications of Cord-Rubber Composite Theory to Tires 162
Obstacle Enveloping Characteristics 162
Tread Wear Resistance 163
Vibrational Characteristics of Radial Tires 167
Stress Analysis 168
Ply Steer Behavior 171
3.1.5. Limitations of Theory/Concluding Remarks 172
3.1.6« Acknowledgements 174
3.1.7. Principal Notation....:. 174
References 175
3.2. Loss Characteristics of Cord Rubber Composites2 179
3.3. Failure Mechanisms of Cord Reinforced Rubber2 1 93

1 Central Research Laboratories. Firestone Tire ind Rubber Co., Akron. Ohio 44317.
2 Contributed by S. K. (lark. Dep'l of Mechanical Engineering, University of Michigan. Ann Arbor, Michigan

123
124 MECHANICS OF PNEUMATIC TIRES

3.1.1 INTRODUCTION AND HISTORICAL


PERSPECTIVE
The stiffness and strength behavior of the constituent cord and rubber
components of the pneumatic tire have been studied in detail by investiga
tors in the industry since the early days of tire production. For example,
the effect of twist on the stress-strain properties of cotton tire cord and the
reinforcing effect of carbon black on the modulus of natural rubber were
well known phenomena many decades ago. Similarly, differences between
the elastic and viscoelastic behavior of cord and rubber have been well
documented since the 1940's. On the other hand, the material properties
(specifically, the elastic constants) of the cord-rubber composite system
that comprises the tire are not as well known and have only begun to re
ceive serious attention in the last decade. These tire elastic properties,
whether they be referred to as stiffnesses, compliances or moduli, are ani-
sotropic—i.e., they vary with direction.
It is the properties of the anisotropic cord-rubber composite that pri
marily control the overall performance characteristics of pneumatic tires.
Yet, much of present day tire design has been dominated by the separate
influences of rubber compounding and textile technology. This situation
has perhaps arisen because the functions of the individual cord and rubber
components of the tire are well known, at least qualitatively: the low mod
ulus, high elongation rubber contains the air and provides abrasion resis
tance and road grip; the high modulus, low elongation cords provide rein
forcement for the rubber and carry most of the loads applied to the tire in
service. However, it is shown later in this discussion that, in order to opti
mize a given tire performance parameter, a knowledge of the combined
cord-rubber composite material properties is required. For example, the
crown angle for maximizing the tread wear resistance for a steel-belted
radial tire is not the same as for a rayon-belted radial of the same size.
While this fact could be empirically established by "trial and error" tire
building and test methods, such a program involving the wide variety of
available tire cord and rubber compounds would be time consuming and
expensive. Thus, composite material mechanics, even though in its early
stages of development as far as cord-rubber systems are concerned, can
and should be used to economically investigate the technical merits of po
tential tire designs.
Historically, the first reference in the scientific literature to what later
came to be called anisotropy was observed in the experiments of Buffon
[1] in 1741 in a study of the load-deflection and rupture properties of
oaken beams. It was not until the 1820's, however, that such behavior
could be accommodated theoretically in the generalized Hooke's law1 con
tained as part of the linear theory of elasticity developed by Cauchy, Pois-
son and Navier. Up to the 1930's, solutions to the equations governing the

1 The generalized Hooke's law for an anisotropic medium is a set of six algebraic equations with the six components of
stress represented as a homogeneous linear function of the six components of strain with 21 independent elastic constants.
While the generalized Hooke's law is not specifically employed herein, familiarity with its various simplifications is helpful
for better understanding the material properties of cord-rubber composites, cf , e.g.. Lekhnitskii [2] and/or Hearmon [3].
CORD REINFORCED RUBBER 125

stress and deformation behavior of anisotropic media were motivated pre


dominantly by problems of crystal physics although civil engineers in lim
ited numbers were beginning to analyze steel reinforced concrete slabs as
anisotropic plates. However, it was with the continued use of wood in gen
eral and plywood in particular as a nonisotropic structural material in the
aircraft industry that scientists and engineers in large numbers entered
and helped to further develop this field. In the tire industry, the first quan
titative study of cord-rubber elastic properties was published in Germany
in 1939 by Martin [4] who analyzed bias ply aircraft tires using thin shell
theory applied to spheres and cylinders to approximate toroidal tire be
havior and (what is referred to herein as) modified netting analysis to pre
dict the orthotropic composite elastic constants (cf. Martin's Fig. 35). A
period of approximately 20-25 years elapsed before the topic of cord-rub
ber properties was again seriously studied—by Clark [5-7] in the USA,
Gough [8] in Great Britain, Akasaka [9] in Japan, and Biderman, et al.
[10] in the Soviet Union. Clark's work represents a culmination of effort
undertaken at the University of Michigan beginning in 1959; Gough's
classical investigations were initiated in 1952-53 but remained unpub
lished until 1967; Akasaka's theoretical and experimental studies, too nu
merous to cite individually, considered the deformation response of vari
ous cord-rubber structures (flat plates, cylinders and toroidal shells);
Soviet research in this area was employed primarily as an aid in studying
critical speed (traveling wave) phenomena in bias ply tires. While all this
work proved useful for studying certain problems of tire mechanics, most
of the equations developed by these investigators could not be used to pre
dict or explain certain important physical phenomena associated with the
deformation characteristics of laminated cord-rubber structures such as
coupling between bending and stretching behavior, the effects of ply
stacking sequence on performance, and interply stress-strain response.
Our present knowledge of the mechanics of composite materials and lami
nated structures is contained in two excellent texts by Ashton, et al. [11]
and Jones [12] written, however, for the study of rigid composites with no
reference to flexible cord-rubber systems.
The objectives of "Elastic Properties of Cord Reinforced Rubber", are:
a. to outline on the phenomenological level the fundamentals of com
posite material mechanics specifically applicable to cord-rubber sys
tems and
b. to indicate how cord-rubber composite material properties can be
used to predict trends in certain tire performance parameters.
These objectives are directed to the analysis of structural laminates (viz.
tires) starting with cord and rubber properties (as diagrammed in Fig. 3.1),
and are not meant to be directed to the determination of tire composite
material properties per se.
Only the static elastic stiffness properties (i.e., "moduli") of cord-rubber
composites are considered in this section; viscoelastic and strength behav
ior, though less amenable to analysis, are considered in Sections 3.2 and
3.3, respectively. Many of the figures and tables contained in Section 3.1
can be found, with appropriate modifications, in a previous paper by Wal
ter, et al. [13] which has essentially been updated in this' study not only to
correct errors but to reflect the tremendous advances that have been made
in our understanding of this field of tire mechanics in the last several
years.
126 MECHANICS OF PNEUMATIC TIRES

Srec'»u-Y GENERALLY
ORTHOTROPIC ORTHOTROPI
RUBBER^ PLY PLV
^" "~J PROPERTIES PRIN. MATERIAL AXES ( W/JW
PROPERTIES /^yvy^
CORD

MULT i -PLY
LAMINATE

FIGURE 3.1. Flow-Chart showing objectives of cord-rubber composite analysis.

3.1.2. CONSTITUENT MATERIAL PROPERTIES


Rubber Elastic Constants
In this survey, rubber will be treated as a homogeneous, isotropic mate
rial with two independent elastic constants. In the structural analysis of
cord-rubber composites which follows, it is convenient to employ the
Young's modulus Er and the Poisson's ratio νr of rubber as the two inde
pendent constants, although in some problems of rubber physics, it is de
sirable to use the rubber shear and bulk moduli (denoted by Gr and K,, re
spectively) as the independent constants. In any case, these four elastic
constants are related to one another, in the limit of small strains, through
the expressions
E, 3AX1 ~
2(1 + 2(1 +
Young's modulus, also commonly referred to as the modulus of elastic
ity, is the initial slope of the rubber stress-strain curve. It is usually deter
mined from a uniaxial tension test but can also be determined from com
pression, bending or torsion experiments. Young's modulus may be as low
as 100 psi for some non rein forced (unfilled) elastomers to as high as
100,000 psi for highly vulcanized (high sulfur) compounds such as ebo
nite. (Rubber elasticity theory, based on molecular considerations, pre
dicts that Er ~ 150-200 psi at room temperature for an unfilled vulcanizate
depending on the degree of cross-Unking [14].) The Young's modulus or
initial slope of the stress-strain curve of a rubber specimen is affected by
physical testing and chemical vulcanization parameters, including (a) the
rate of deformation and temperature of the test (that is, the viscoelastic na
ture of the rubber), (b) the cyclic loading history of the specimen (that is,
appreciable stress-softening occurs in filled systems), (c) the compounding
CORD REINFORCED RUBBER 127

ingredients mixed with the rubber, such as carbon black, sulfur, and oil,
and (d) the state of cure of the rubber. In fact, as a further refinement, the
carbon black-rubber system which we are treating as a homogeneous, iso-
tropic continuum can be considered as a two-phase paniculate composite
consisting of rigid inclusions (carbon black particles) in a compliant ma
trix (rubber). These factors, and their ramifications, are discussed in detail
by Ferry [15], Kraus [16], and Nielsen [17] and are not further considered
herein.
The so-called 300% modulus, used for quality control purposes in the
rubber industry, may bear some, little or no relationship to the Young's
modulus; it, of course, is not a modulus at all in the mechanics usage of
the word but rather is the stress (with the same dimensions as modulus)
required to produce 300% elongation in a uniaxial tension test. Exten-
sional strains of 300% or greater are rarely if ever encountered in any of
the tire components under service conditions (except in the tread region
during the abrasion process)—i.e., the 300% modulus is not required for
characterizing the elastic constants of cord-rubber composites.
Poisson's ratio, in a small strain uniaxial tension test, is the absolute
value of the ratio of lateral contraction to longitudinal extension of the
loaded rubber specimen. Poisson [18] at one time presented erroneous ar
guments based on the molecular theory of elastic bodies developed by
Navier that ν = 0.25 for all elastic materials, but actual measurements
show that this ratio varies between 0.3 and 0.4 for most metals and plas
tics, and approaches 0.5 for elastomers. Hence, when rubber is deformed,
little if any volume change occurs compared to that which occurs in other
materials, so that:
v, -» 'A, K, -> oo, and E, -» 3G,.
In other words, only one independent elastic constant exists for in
compressible isotropic materials.
The large differences that exist among typical (approximate) values of
the four elastic constants (E, G, K and ν) of various materials contrasted to
rubber (without carbon black reinforcement) are apparent from inspection
of Table 3.1.

TABLE 3. 1 Typical Values of the Four, Interrelated Elastic Constantsfor Various


Homogeneous, Isotropic Materials
Young's Modulus Shear Modulus Bulk Modulus Poisson's Ratio
Material E, psi G, psi K, psi r
Steel 29 x 10* 11 x 10* 24X10" 0.30
Glass 8.7 x 106 3.5 X 10* 5.4 X 106 0.23
Lead 2.3 X 106 0.8 X 10* 1.1 X 10" 0.43
Epoxy 0.5 X 106 0.2 X 10* 0.5 X 106 0.35
Rubber0 ISO 50 0.3 X 10* -0.50
• Nonreinforced

Implicit in the preceding discussion of rubber elastic constants is the as


sumption that the Young's modulus and Poisson's ratio are identical in
tension and compression. If experimental results indicate that the rubber
compound in question behaves bilinearly with a discontinuity at the origin
of the stress-strain curve, then the following expression relating elastic
constants is applicable [19]
128 MECHANICS OF PNEUMATIC TIRES

•12

FIGURE 3.2. Illustration ofplane stress state for isotropic material (Eq. I).

(1 + <) + (! +
where the superscripts + and - denote tensile and compressive properties,
respectively. When E? = E~, and v* — v~, the familiar expression given
previously is recovered.
The elastic constants E, and νr are contained as pan of the Hooke's law
for isotropic materials which relates in-plane stresses (a,, σ2, τ12) to in-
plane strains (e,, €2, y,2), under so-called plane stress conditions (lateral di
mensions large compared to thickness. Fig 3.2), that is,

e, = — (a, - vra2)

(3.1)

The relationships of Eq. 3.1 are obtained from the three dimensional stress-
strain equations (Hooke's law) of the theory of elasticity by assuming that
the normal stress a, and the shearing stresses τ13 and T.,, are negligibly
small compared to 0,, σ2 and τ12 [20]. An analogous set of stress-strain rela
tions can be written for the case of "plane strain" governing the deforma
tion behavior of long cylindrical bodies; since tire geometry and loading
conditions more closely approximate plane stress rather than plane strain
conditions, equations for this latter case are not considered. Note that re
gardless of the particular form of Hooke's law, two independent elastic
constants characterize isotropy.
In many of the numerical examples employed in Section 3.1, values of
the rubber elastic constants given in Table 3.2 are used; the rubber com-

' Tbex quantities (n, rp, rJ3) are the imerply stress component! dacmied in Section 3.1.3. under the beading "Multi-
Ply Systems: Inleilammar Deformations"
CORD REINFORCED RUBBER 129

TABLE 3.2 Typical values ofelastic constantsfor rubber used in calendered plies oftires

Young's Modulus Poisson's Ratio


Rubber Skim Stock I , . psi r,
Textile Body Ply 800 O49
Textile Tread Ply 3000 0.49
Steel Tread Ply 2000 0.49

pounds listed are those which encapsulate cord to form a calendered ply
(so-called skim stocks). The values reported were obtained from slow
speed, room temperature, uniaxial tension tests. It is apparent that the
compounding ingredients used in these (and other) tire rubbers have a
pronounced effect on increasing the Young's modulus above that associ
ated with nonreinforced compounds (~ 1 50 psi).
Cord Elastic Constants
The hundreds of continuous, oriented, polymeric filaments that consti
tute the typical organic plied and cabled cord used in tires should individ
ually be considered as being transversely isotropic with five independent
elastic constants: (a) an extensional Young's modulus, (b) an extensional
Poisson's ratio, (c) a transverse Young's modulus, (d) a transverse Pois
son's ratio, and (e) a torsional (shear) modulus. For such filaments, isot-
ropy exists in planes perpendicular to the direction of fiber drawing as dis
cussed by Ward [21]. It is apparent then that twisted tire cord, which also
has a certain amount of void content between its filaments and plies, can
not be considered as a homogeneous, isotropic material like rubber.
Brewer [22] in his stress analysis of bias ply aircraft tires, approximated
the material properties of twisted cord as being transversely isotropic
which is strictly true only for a single fiber without twist—i.e., mono-
filament. However, to be consistent with approximations we employ later
in this section (Eqs. 2 to 6), we require only the extensional Young's mod
ulus and Poisson's ratio as well as the shear modulus of the cord; we will
neglect the transverse Young's modulus and Poisson's ratio of the cord for
which no theoretical or experimental data are available. It appears that
the transverse Young's modulus and Poisson's ratio of the cord do not
strongly affect many of the properties of two-dimensional cord-rubber
composite systems as used in tires. Whitney [23] has shown, however, that
transverse isotropy of filaments can have a significant effect on the elastic
constants of rigid graphite-epoxy composites.
In addition to the complications introduced by viscoelastic effects
(strain rate and temperature dependent material properties), tire cord elas
tic constants are also significantly affected by the amount of ply yarn and
cable twist employed. (Twist is given in turns per inch. With steel tire
cord, the term "lay", which is the reciprocal of the twist, is preferred.) For
example, to a first approximation, the Young's modulus Ec of twisted sin
gles yarn in tension decreases with increasing twist according to the ex
pression [24]

where Ef is the filament modulus, R the yarn radius, and T the twist; the
Poisson's ratio v, of two ply cord in compression decreases with increas
ing twist according to the relation νc = (4π2R2T2)-1 [25].
130 MECHANICS OF PNEUMATIC TIRES

In some cases, twisting fiber into tire cord can result in as much as a
one-third decrease in the extensional Young's modulus for typical belt ply
cords and a one-half decrease in the extensional Young's modulus for typ
ical body ply cords. However, twisting is needed in order to provide ade
quate cord fatigue life under tire service conditions. Cord structure is dis
cussed by Hearle, et al. [26] from a textile technologist's point of view,
while cord fatigue phenomena, cord-to-rubber bonding problems, etc. are
discussed by Takeyama and Matsui [27]. Some idea of the range of values
for the Young's modulus of typical belt and body ply cords can be gained
from inspection of Table 3.3. '
The stress-strain properties, including Young's moduli, of a variety of
belt cord materials are contained in the data of Draves, et al. [28]. The
shear moduli of organic tire cord constructions are taken to be approxi
mately 700 psi [22], though measurements to confirm this value are lack
ing; measurements on steel cord (5 × 1 × 0.010 in) are more easily con
ducted and indicate that Gc is approximately 1 × 106 psi [29]. Poisson's
ratios of tire cords are often in excess of 0.5 due to twist (the higher the
twist level, the larger νc is).

3.1.3. Cord-Rubber Composite Properties


One-Ply Systems: Engineering Elastic Constants
A unidirectional, calendered ply of cord and rubber is depicted in Fig.
3.3. Such a system, which is n on homogeneous, anisotropic, and more diffi
cult to characterize than cord or rubber alone, is termed "specially ortho-
tropic" [1 1]. It belongs to that class of material known as filamentary com
posites. The cord direction, denoted by "1", and the direction
perpendicular to the cord, denoted by "2", form an axis system referred to
as the natural or principal material directions of the ply. Five engineering
elastic constants, four of which are independent, are required to define the
material properties of the unidirectional cord-rubber system: the longitu
dinal Young's modulus £„ the transverse Young's modulus E2, the in-
plane shear modulus G12, the major Poisson's ratio i>12, and the secondary
Poisson's ratio ν21. These are based on the overall ply thickness dimension,
just as in any isotropic material.
As might be expected, such composite elastic constants primarily de
pend on the differing elastic properties and concentrations of the cord and
rubber. The constants can be determined experimentally from direct mea

TABLE 3.3 Typical values of Young's Modulusfor tire cord


Young's Modulus
Cord Construction Ec> psi
Belt Ply
5 x 1 X 0.010 in. Steel 15.9 X 10*
1500/2 Kevlar 3.6 X 10*
1650/3 Rayon" 1.6 X 106
Body Ply
1000/2 Polyester 575,000
840/2 Nylon 500,000
- This is a relatively low twist rayon (with a high Young's modulus) compared to the
higher twist level rayon used in the body plies of tires.
CORD REINFORCED RUBBER 131

FIGURE 3.3. Unidirectional calendered ply of cord and rubber showing natural or principal
material axes.

surements made on the calendered ply or can be approximated from the


properties of the individual cord and rubber components of the ply using
the following equations:
(3.2)
v,(\ - vc) (3.3)
- vc)]
••'•" Ec(\ - ve) + f, E,(\

r"

(3.6)
where f,, f2 are factors depending on cord geometry and spacing, and v,. is
the volume fraction of the cord in the calendered ply given by

with R = cord radius, t = calendered ply thickness, and epi = cord end
count (ends per inch). Note that for a planar network of parallel cords as
used in the plies of tires, cord volume fraction is identical to the ratio of
cord cross-sectional area to the total area of cord and rubber.
Equations 3.2-3.5 are usually referred to as the Halpin-Tsai equations
[11, 12, 30]. Equation 3.6 expresses the well-known reciprocity relation of
structural mechanics; it of course indicates that of the five elastic constants
for the specially orthotropic ply, only four are independent (G12 plus any
three of £„ E2, ν12 and ν21). Note that, physically, the Poisson's ratio ν21 is
associated with a contraction in the 2 direction caused by a tensile stress in
the 1 direction (i.e., i>,2 = -∊2/∊1); «*21 is defined in a similar and opposite
manner.
For cords of nominally circular cross-section, as is generally the case for
tire reinforcements, and for Ec » E, as is generally the case with cord-rub
ber composites, f, = 2 and f2 = 1. Thus Eqs. 3.4 and 3.5 for predicting the
transverse and shear moduli of the unidirectional composite reduce to
those previously employed by Walter, et al. [13]—i.e.
132 MECHANICS OF PNEUMATIC TIRES

"12 r/~ _i_


[Gc + /- r1 - /~\.,
G,- (Ge Gr)ve]i
If the tire cord cross-section is non-circular such as with ribbon or tape-
like cord [31], Eqs. 3.4 and 3.5 should be employed with appropriate val
ues of f, and fc [1 1, 12].
Equations predicting the time-dependent moduli of filamentary com
posites—both static [32] and dynamic [33]—have been developed by
Hashin based on the correspondence principle of viscoelasticity. The Hal-
pin-Tsai equations, e.g., can be easily extended to embrace such behavior
for the case of purely elastic fibers embedded in a matrix possessing creep-
relaxation characteristics [30]—i.e., the viscoelastic response of a uni
directional composite can be estimated if the tensile and shear relaxation
moduli of the rubber are substituted for the tensile (£,) and shear ((/,)
elastic moduli. This estimate is probably valid for steel cord-rubber sys
tems but experimental corroboration is still lacking. Viscoelastic and loss
characteristics of composites are discussed further in Section 3.2.
The derivation of Eqs. 3.2-3.5 is associated with that part of the science
of composite materials known as "micromechanics"—i.e., the study of
one-ply composite behavior which recognizes the inhomogeneous nature
of the ply by utilizing properties and concentrations of cord and rubber.
On the other hand, the study of multi-ply composites (i.e., laminates)
which ab initio assumes each ply of the system to be homogeneous but an-
isotropic with the effects of cord and rubber detected only as avearge (or
"smeared") properties is termed "macromechanics".
While the various stress-strain equations employed later in this section
for analyzing tire behavior are based on the assumptions of macrome
chanics, it is worthwhile to examine in a cursory fashion some of the ele
mentary problems of micromechanics. Equations 3.2 and 3.3 specifically
referred to as the rule or law of mixtures, are illustrative. The longitudinal
modulus /;, can be determined from a mechanics-of-materials approach
by considering the cord and rubber to act as springs (Hookean solids) in
parallel when a load is applied to the composite coincident with the cord
direction as shown in Fig. 3.4. In this case the relations
e, = ec = €„ oc = £cc,, a, = E^
with the definition of cord volume fraction lead directly to Eq. 3.2. Equa
tion 3.5 for the major Poisson's ratio r,., is derived in a similar manner.
An expression for the transverse modulus E2 can be obtained by consid
ering the cord and rubber to act as springs in series when a uniaxial load is
applied to the composite perpendicular to the direction of the cords as
shown in Fig. 3.5. In this case the relations
o2 = ae = an ec = Oi/Ea er = a-JE,
with the definition of cord volume fraction lead to
1 vc l-vc _
£,v, (3'7)
which differs only slightly from Eq. 3.4 in the range of cord volume frac-
CORD REINFORCED RUBBER 133

LOAD

tCORD ^RUBBER

fLOAD

CORD-RUBBER COMPOSITE PARALLEL MODEL

FIGURE 3.4 Hookean model for calculating longitudinal modulus Et of calendered ply of
cord and rubber.

tions and constituent moduli used in cord-rubber systems. For anisotropic


reinforcements, Whitney [23] recommends using the transverse modulus
of the cord rather than the longitudinal modulus; however, such properties
are generally unavailable for twisted tire cords.
Equation 3.2 for predicting the longitudinal composite modulus with
equivalent strains assumed in the cord and rubber is analogous to the clas
sical Reuss treatment. While the elementary micromechanical analyses
described above were utilized to obtain mechanical properties of filament
ary composites, similar reasoning can be applied to obtain such properties
for particulate composites (e.g., carbon black-rubber systems) and semi-
crystalline polymers [34]. Further, the same approach can be used to de
rive equations for predicting the thermal, electrical and other "transport"
properties of filamentary (and particulate) composite systems [35].
It should be recognized that Eqs. 3.2 and 3.7 are special cases of the
general empirical relation
£" = £>r + £*l-vc) (3.8)
where n is an exponent and £ represents either the longitudinal (n = 1) or
transverse (n = - 1) modulus of the ply. While it is known that the parallel
(n =" 1) and series (n = — I) models rigorously represent upper and lower

LOAD LOAD CORD RUBBER

CORD-RUBBER COMPOSITE SERIES MODEL

FIGURE 3.5 Hookean modelfor calculating transverse modulus E2 of calendered ply of cord
and rubber.
134 MECHANICS OF PNEUMATIC TIRES

10

UJ

0. 0.2 0.4 0.6 0.8 1.0


CORD VOLUME FRACTION vc
FIGURE 3.6 Dependence of cord-rubber composite modulus on volume fraction for 1000/2
polyester-rubber ply.

bounds, respectively, for the actual modulus of composite materials [12],


most experimentally determined properties fall between these two ex
tremes corresponding to values of n other than ±1, i.e. -1 < n < 1.
Equation 3.8 with a logarithmic scale for the ordinate is plotted as a
function of volume fraction in Fig. 3.6 along with Eq. 3.4a for a typical
1000/2 polyester-rubber composite used in radial tire body plies (Er =
1060 psi, Ec = 575,000 psi, νr = 0.49 and νc = 0.66). While Fig. 3.6 covers
the entire range of possible volume fractions (0 to 1), it is important to
note that cord-rubber plies as used in tires usually fall in the range νc =
0.10 to 0.40—i.e., volume fractions are relatively low compared to those
employed in rigid filamentary composites. This apparently low volume
fraction arises because a certain amount of rubber insulation is needed be
tween cords in the same ply (intraply) and between cords in adjacent plies
(interply) to provide against friction a 1 rubbing of the cords and to promote
adquate fatigue resistance of the tire under service conditions. Further, de
pending on the analysis to be conducted, one or more plies of a tire may
be considered in different but equivalent fashions for volume fraction pur
poses. Fig. 3.7 is illustrative; here a two-ply cord-rubber system with each
ply (and the system) having a volume fraction of irR2(epi)/t may be al-
terntively characterized as an equivalent five-ply system with three iso-
tropic rubber plies of different thicknesses (νc = 0) and two higher volume
fraction cord layers of thickness R with »', = irR(epi).
Another volume fraction anomaly occurs in steel cord-rubber systems
due to the relatively open structure of the cord cross-section. Fig. 3.8
shows cord cross-sections for two types of steel cord used in the belts of
radial passenger tires compared to a typical organic tire cord material used
in the body plies. While the nominal diameter of organic cord construc
tions (nylon, polyester, Kevlar, etc.) composed of hundreds of individual
filaments can be used to compute accurate volume fractions, actual cord
area based on individual filament diameters should be employed in the
case of steel.
CORD REINFORCED RUBBER 135

CORD RUBBER

±
O O O O O O O
O O O O O O O
NOMINAL TWO-PLY
(TT o n o n o o
(TT o o u o o n
EQUIVALENT FIVE-PLY
I2'
FIGURE 3.7 Different representations for a nominal two-ply cord-rubber system.

A precaution to be aware of in the application of the Halpin-Tsai or


similar equations to cord-rubber composites is that the initial or Young's
moduli of many textile tire cords are significantly different in the vulcan
ized tire from that in the greige, untreated condition (on which cord tests
are likely to be conducted) due to sensitivity to heat and moisture. When
ever any doubt exists concerning the suitability of these equations for pre
dicting the elastic constants for a one-ply system, it is advisable to make
direct experimental measurements on the ply. The various test procedures
necessary for this experimental determination of the five elastic constants
of a unidirectional composite are described by Jones [12], Bert [36] and
Calcote [37].
Some measure of the validity of Eqs. 3.2 and 3.7 for predicting the lon
gitudinal (E1) and transverse (E2) moduli of cord-rubber composites is
shown in Figs. 3.9a-9b. comparing theory and experiment [38] for experi
mental 1500/2 Kevlar-rubber plies with different cord volume fractions
achieved by varying cord end count. It appears that the longitudinal mod
ulus is adequately predicted, while the results for the transverse modulus
contain more uncertainities over this range of cord volume fractions. Since
the simple rule of mixtures relations (viz., Eqs. 3.2 and 3.7) do not explic
itly contain terms associated with the degree of adhesion between cord
and rubber, cord geometry or cord spacing, it is reasonable to conclude
that these factors do not appreciably influence the longitudinal modulus.
In contrast, the transverse modulus (and the shear modulus) of a uni
directional cord-rubber ply is probably appreciably influenced by these
factors. While the transverse and shear moduli of cord-rubber composites
are greater than that of rubber alone, the actual values measured strongly
depend on the behavior of the rubber interacting with the cord through
the adhesive interface.
More advanced concepts than mechanics of materials must be em
ployed (i.e., theory of elasticity) to establish expressions more precise than

STEEL STEEL POLYESTER

1x2+7x0.009 IN. 1x5x0.010 IN. 1000/3

FIGURE 3.8 Typical tire cord cross-sections.


136 MECHANICS OF PNEUMATIC TIRES

300,000"
M
a.

UJ
en 200,000 Eq. 2
3 • EXPERIMENT

< 100,000-
z
o Ec = 875,000 psi
Er = 1150 psi
o

.10 .20
CORD VOLUME FRACTION vc
FIGURE 3.9a Comparison between theoretically predicted and experimentally measured
values of longitudinal modulus E, as a Junction of volume fraction vcfor 1500/2
Kevlar-rubber ply.

~ 3000
M
Q.

CM
UJ

2000
O
o
LJ
CO 1000
(E
UJ
Eq. 4a
en Eq. 7 Ec =875,000 psi
EXPERIMENT Er = M50 psi

.10 .20 .30


CORD VOLUME FRACTION vc

FIGURE 3.9b Comparison between theoretically predicted and experimentally measured


values of transverse modulus E2 as a function of volume fraction vcfor 1500/2
Kevlar-rubber ply.
CORD REINFORCED RUBBER 137

those given by the simple rule of mixtures for the various elastic constants
of a unidirectional composite. The Halpin-Tsai relations (Eqs. 3.2-5) are
semi-empirical in nature (viz. Eqs. 3.4 and 3.5) but are based on more
complicated micromechanical analysis. While the predictions of these
equations have been shown to be in reasonable agreement with measured
values of the mechanical properties of filamentary, unidirectional com
posites, it is recommended that the elastic constants be experimentally de
termined for use as "input" to the equations for predicting laminate prop
erties unless such measured data are not available.
The five elastic constants (£„ E2, G12, V12, V21) appear in the Hooke's law
for orthotropic materials which relates in-plane stresses (σ1, σ2, τ,2) to in-
plane strains (€„ ∊2, γ12)—again under plane stress conditions (Fig. 3.10),
that is,

a*
(3.9)
E2 Bt

Yl2 =

Equation 3.9 for a specially orthotropic ply of cord and rubber is only
slightly more complicated than Eq. 3.1 for an isotropic rubber sheet.
As an example of the use of these relationships, consider a typical 1650/
3 rayon-rubber ply used in pairs in the body of an HR78-15 radial tire
with a green (uncured) end count of 22 epi. In the cured tire, the end count
varies continuously from a maximum at the bead (22 epi) to a minimum at
the crown (14 epi). We shall make our calculations at the sidewall where
the cured end count is approximately 18 epi and the cured ply gage is
about 0.048 in. Using Er = 800 psi, Ec = 740,000 psi, νr = 0.49, νc = 0.66,
and νc = 0.34, we obtain for the cord-rubber composite properties (using
Eqs. 2 to 6 with {, = 2 and & = 1);

•12

•12

FIGURE 3.10 Illustration ofplane stress for specially orthotropic material (Eq. 3.9).
138 MECHANICS OF PNEUMATIC TIRES

£, = 252,500 psi
£2 = 2,040 psi
Gn = 363 psi
v,2 = 0.547
v21 = 0.004.
Thus, with such material properties known from calculation (or measure
ment), radial tire sidewall strains, e.g., can be obtained from the Hooke's
law (Eq. 3.9) if the stresses can be calculated for some particular loading
condition. Note that while the stress and strain level in the radial tire side-
wall depends on the number of plies (or total stiffness), the elastic con
stants (/;„ E2, G12, V12 and »'.-,) of the unidirectional sidewall plies do not.
In order to study an area of a tire more structurally complex than a
radial tire sidewall (where specially orthotropic theory applies), the "off-
axis" response of a calendered ply must be known. In an off-axis test of a
single ply, the direction of the applied load is not coincident with the natu
ral or principal material axes of the system. Such a system is termed "gen
erally orthotropic" [11]. Its response to load is considerably more com
plicated than that of the unidirectional specially orthotropic ply since an
applied tensile stress now produces shearing strain as well as the expected
normal strains.
The strain-stress relations governing the off-axis response of a single ply
of cord and rubber can be written (Fig. 3.1 1) as [37]:

y * \
f ^ —£—
*ji r> — "xy
v - j? — AT
f^y'xy (3.10)

•xy

uxy

FIGURE 3.1 1 Illustration of off-axis stress application (x, y directions) to a generally


orthotropic ply with principal material axes in the I, 2 directions (Eq. JO).
CORD REINFORCED RUBBER 139

where the properties in the off-axis directions (denoted by subscripts x and


y) are related to the natural or principal material axes (denoted by the sub
scripts 1 and 2) by
1 cos40 ,' 'f 1 ' 2rI2]' sm26cos26
. .„ .a + —=—
sin40

sin40 " 1 2x12~ . , cos'0

(3.n)
+ "21
£,
^., 1 ( 1 2c1
sm20 - ----r
[_ £2 £, 2 [G,2 £,
cos20 sin2^ 1 r 1 2»-12]
_ -JF---s-
£2 £, - T
2 1[GI2
r---F^
£, J
The complexity of Eqs. 3.10 and 3.11, which describe the off-axis re
sponse of a single ply of cord and rubber, compared to Eq. 3.1 for the iso-
tropic rubber sheet and Eq^ 3.9 for the specially orthotropic cord-rubber
sheet, is apparent. The coefficients λx and λy which couple normal stresses
with shearing strain (and shearing stress with normal strains), vanish in
the case of isotropy and special orthotropy. Thus, for example, if a ply of
cord and rubber is loaded by a uniform shearing stress τxy = τ with a, *• σy
= 0, the strain components are given by
e, = —AXT, €y = -Ayr, y,, = r/G
so that, in general, the ply contracts (or elongates) in addition to shearing.
As an application of the equations which govern the off-axis response of
a generally orthotropic system, consider a typical 840/2 nylon-rubber
composite used in four-ply bias tires with the cord at 32 epi and the ply
gage at 0.027 in. Figures 3.12a and 3.12b show the variation of the elastic
constants (Ex, Gxy, Vxy and vyx) with cord angle θ from 0° to 90°. The cord
angle in a flat pad uniaxial tension test is the angle between the cord and
the applied load. Note that

v,*(ff) - v^fi-n - 0)
and when θ = 0° and 90°, the off-axis constants reduce to the specially or
thotropic properties. For any particular cord angle, strains can be calcu
lated for a given state of stress using Eq. 3.10.
Some measure of the degree of anisotropy of a unidirectional composite
can be established from the value of the ratio of longitudinal to transverse
modulus (E1/E2) of the ply. For an isotropic material, this ratio is unity.
Typical results for rigid and flexible composites are listed in Table 3.4.
Another parameter that governs the response of a structure to load is
the magnitude of the ratio of Young's modulus to shear modulus. This ra-
140 MECHANICS OF PNEUMATIC TIRES

Ec « 500,000 psi
ER - 800 psi

'Cfr 30° 60C 90°


CORD ANGLE 9
FIGURE 3. 1 2a Variation of Young's modulus Ex and shear modulus Gxy with cord angle 6for
840/2 nylon-rubber ply.

tio is again very high for cord-rubber composites compared to both rigid
composites and homogeneous, isotropic materials, which indicates that
shear deformation, and the increased flexibility associated with it, can be
important in beams, plates and shells made with cord and rubber com
ponents. Thus, material properties (E/G ratio), in addition to the expected

EC =500,000 PSI I/C = .45


Er =800 PSI vr = .49
£•1.0
A
QC
Q O __^X ^
^ •" ""N

o -6
H
* .4

z
o
V)
o
Q-
0° 30° 60" 90°
CORD ANGLE 9
FIGURE 3.12b Variation of Poisson's ratio vxy and vyl with cord angle 6 for 840/2 nylon
rubber ply.
CORD REINFORCED RUBBER 141

TABLE 3.4 Longitudinal and transverse moduli comparison for rigid andflexible composites
Longitudinal Transverse Degree of
Filamentary Young's Modulus Young's Modulus Anisotropy
Composite System E,, psi Ej, psi E,/E2
Glass-Epoxy 7,500,000 2,600,000 2.9
Graphite-Epoxy 30,000,000 • 750,000 40
Nylon-Rubber 163,000 2,000 80
Rayon-Rubber 253,000 2,000 125
Steel-Rubber 2,540,000 3,000 850

geometrical factor (depth-to-span ratio), must be taken into account in the


design of such structures. Tarnopol'skii, et al. [39] provide a good review
of this topic with applications to beams.
One-Ply Systems: Compliance and Stiffness Constants
The preceding development dealing with composite properties with em
phasis on the engineering elastic constants gives the tire designer some
physical insight into the effects on mechanical properties of combining
two dissimilar materials—cord and rubber—into one system. For ex
ample, the individual steel mononlaments used in the belt cords of a radial
tire initially possesses a Young's modulus of 28 to 30 X 106 psi; when five
of these filaments are twisted into a 5 X 1 X 0.010 in. cord construction,
the modulus decreases to 15 to 18 × 106 psi; when this steel cord is com
bined with rubber (with a modulus of 800 to 3000 psi), the composite
modulus in the direction of reinforcement is further reduced to 1 to 3 ×
106 psi depending on cord end count and ply thickness; and then this value
is achieved only if the composite system is loaded in tension. Compressive
Young's moduli for steel cord-rubber composites may be three orders of
magnitude lower.
In order to develop concisely and expeditiously the appropriate repre
sentation of composite elastic constants for laminated or multi-plied sys
tems, it is convenient to work with a more mathematical and less physical
form of Hooke's laws for both the specially orthotropic and off-axis single
ply. Thus, the Hooke's law relating strains to stresses in matrix notation
for specially orthotropic composites loaded in a plane stress state equiva
lent to Eq. 3.9 is given by

6, 5,, 5,2 0

d 512 522 0 (3.12)


Yl2j 0 0 5«

where [S] or Sv represents the compliance matrix, the elements of which


are symmetric about the main diagonal and which may be interpreted as
reciprocal moduli.
142 MECHANICS OF PNEUMATIC TIRES

In abbreviated matrix notation these relations are written


{£}=[$] {a}.
Equation 3.12 may be inverted to obtain

"c,, 2,2 0 " 6,

3.2 222 0 €2 (3.13)


TI2. 0 0 e« Y.2

where [Q] or Q,, represents the so-called reduced stiffness matrix. The en
gineering constants, the elements of the compliance matrix and the ele
ments of the reduced stiffness matrix are easily related to one another and
the interrelations are given for ready reference in Table 3.5.
In order to now express stresses (or strains) in the natural or principal
1,2 material axes in terms of an arbitrary x,y coordinate system (Fig. 3.1 1),
it is only necessary to employ the equations governing the transformation
of second order tensors routinely used in mechanics of materials (Mohr's
circle) and in dynamics (moment-of-inertia calculations), i.e.
0, COS20 sin20 2 sin 8 cos 0
»2d
sin'ff cos20 -2 sin 0 cos 8
«a (3.14)
TU -sin 0 cos 0 sin 8 cos 0 cos20 — sin20

TABLE 3.5 Interrelations among the differentforms ofthe elastic constants


Engineering Constants
E, £2 "|2 "21 G|2

Compliances 1
s,,
Stiffnesses Q..Q22-Q.22 Q||Q22-Q|22 Ql2 Ql2 Q
0.22 Oil 022 Oil

Compliances
s,, Sc
22 I -'
c
^66

Stiffnesses 022 On Q.2 i


Q||Q22-Q|22 QllQ22~Q.2 QllQ22~Ql2 Q«

Engineering 1 1 "12 1
Constants E, Ej E, G,2

Stiffnesses
Q,, 022 Ql2 Q«6

Engineering E, £2 "12 E2 p
Constants 1 - "12 "21 I-"|2"2I 1-".2"2I

S22 SM -SI2 1
CORD REINFORCED RUBBER 143

or as is often written

[T]

where the coordinate transformation matrix [T] is given by

cos20 sin20 2 sin 0 cos B


m sin20 cos20 -2 sin 0 cos 0 (3.15)
—sin 6 cos 0 sin 6 cos 0 cos20 — sin20

Note that [T] is not symmetric in contrast to all of the matrices employed
herein for characterizing material properties.
Inversion of Eq. (3.14) produces

(3.16)

where the superscript -1 denotes the matrix inverse (in this special case
replace θ by -θ in the definition of [T] to obtain [T]~\
In the same manner
e,
«t (3.17)

where the factor Vi is introduced to multiply shearing strain so that tensor


manipulations are valid (since tensor shearing strain equals one-half of en
gineering shearing strain).
Rewriting Eq. 13 with the above in mind produces

c., 0

«2 0 [0 (3.18)
fa 0 0

Substituting Eqs. 17 into 18 and the result into Eq. 16 leads to


t.
= [TTl [Q] [T] e, (3.19)

where the transformed reduced stiffness matrix [Q] is a fully populated 3


X 3 symmetric matrix, i.e.
144 MECHANICS OF PNEUMATIC TIRES

fin fi,2 fi,«

fil2 fi22 fi*

fil6 $26 fi<*

and the Q,, are related to the reduced stiffnesses Q,, as follows:
fin = (?ucos40 + 2(g,2 + 2<2«)sin20cos20 + Q22sin46
fi,2 = (Gn + &2 ~ 4£?«,)sin20cos20 + (?12(sin40 + cos"0)
fiM = Qnsin4B + 2(2,2 + 2266)sin20cos20 -I- Q22cos46 (3.20)
fiia = (Qn ~ Qn - 2g«1)sin0cos30 + (Q12 - Q22 + 2£>66)sin30cos0

fi« = (di + &2 - 2QI2 - 2 cos'0)


In similar fashion, inversion of Eq. 3.19 produces

• «, = (fir *, = [5] CT, (3.2i)


I*? r*y) T*y.

where the transformed compliances Sv are given in terms of the (ordinary)


compliances Sij as follows:

•S,2 = 5,2(sin40 + cos40) + (5,, + 5i2 - 5«)sin20cos20


1X22 •* 5,,sin40 + (25,2 + 56«)sin20cos20 + 522cos40 (3.22)
5|6 = (25,, — 25,2 ~ Sftsjsintfcos3^ — (2522 ~ 2512 ~ 566)sin30cos0
S2t, = (25, , - 25,2 - 5«)sin30cos0 - (2522 - 25,2 - 566)sin0cos30
5^ = 2(25,, + 25i2 - 4512 - 5«,)sin20cos20 + 5«(sin40 + cos4ff)
Recall that the interrelationships among the 5,,, Q,, and engineering con
stants are given in Table 3.5.
Experimentalists generally find it more convenient to work with the
strain-stress (∊-σ) form of Hooke's law as exemplified by Eqs. 3.9 (or Eq.
3.12) for specially orthotropic systems and Eqs. 3.10 (or Eq. 3.21) for gen
erally orthotropic systems rather than the stress-strain (σ-∊) forms. This is
because in experimental procedures it is easier to apply a simple stress of a
given type (e.g., a tensile stress or a shearing stress), and measure the cor
responding strains, e.g., from Eq. 3.21
• i'C /. X C » 4. 0 ~
t* JV"J|I«* ~ "I2°>' ~ 'J|6T*y

The transformed compliance constants S11, S12, and 5,,, can be found by
separately applying the stress σ x, σy and τx y and measuring the strain ∊x in
each case.
On the other hand, for the development of the appropriate equations
governing the response of laminated or multi-ply composites, analysts
generally find it more convenient to work with the σ-∊ form of Hooke's
CORD REINFORCED RUBBER US

law as exemplified by Eq. 3.13 for specially orthotropic systems and Eq.
3.19 for generally orthotropic systems. This will be apparent in the dis
cussion of multi-ply systems which follows.
In any case, whichever form of Hooke's law is employed (a-e or e-a),
there are three different but equivalent sets of elastic constants which can
be used to linearly relate stresses to strains—compliances, stiffnesses or en
gineering constants. However, even in the case of an off-axis (generally or-
thotropic) ply, only four independent constants (which vary with cord
angle ff) are needed for Hooke's law when plane stress geometry and load
ing conditions exist. Variations in these different properties (engineering,
compliance and stiffness constants) with cord angle θ are shown in Figs.
3.13a to 3.13c for a 1000/2 polyester-rubber ply (based on data of Patel, et
al. [40]).
An important consideration to be aware of in interpreting experimental
data for the elastic constants of off-axis tensile test specimens is the large
difference that exists between the modulus Ex and the stiffness (),, as
shown in Fig. 3.14 for a polyester-rubber ply (which is a replot of the
curves contained in Figs. 3.13a and 3.13c). This is important because the
geometry of the test specimen (specifically the length-to-width ratio) and
the manner in which it deforms at the grips govern whether the appropri
ate stress-strain equation is (see Jones [12], p. 67):
a, = E,f, or a, - £,,ex.
Finally, it is commonly observed for filamentary composite materials
that different values of the Poisson compliance Sn are measured depend
ing on whether the unidirectional ply is loaded longitudinally or trans
versely—i.e., the reciprocity relation, Eq. 3.6, is not satisfied. Two proce
dures for accommodating this observation in analysis have been suggested
[41]: (a) rejection of the symmetry hypothesis used in the various elastic
constant matrices (e.g., S12 ≠ S21) and (b) the use of two sets of symmetric
matrices to represent the various elastic constants (one set of properties
corresponding to cords loaded in tension, another different set correspond
ing to cords loaded in compression). Rejection of material symmetry—i.e.,

& ^.6
&
K
UJ
0*1500. O CO
^ 1200. .4 I04 _
or
3 900. I03 O
o CO o
0 600. ~z I02
o CO
to
< 300. I01 ~e>
z
1 0. 0. 10° o
CO '0° 30° 60° 90'
CORD ANGLE 9
FIGURE 3.13a Variation of engineering constants E,, Gfy and v,y with cord angle 8 for
J 000/2 polyester-rubber ply.
146 MECHANICS OF PNEUMATIC TIRES

ICO

0* 30° 60° 90°


CORD ANGLE B
FIGURE 3.13b Variation of transformed compliance constants S,,, SI2, S,6 and SM with
cord angle 6 for 1000/2 polyester-rubber ply.

allowing the compliance matrix to be unsymmetric but vary smoothly with


angular orientation—has widespread implications of a negative nature in
stress analysis since the concept of strain energy is based on symmetry.
Thus, while both models can be used to represent material behavior well,
the second approach based on the use of bilinear symmetric elastic proper
ties is more philosophically desirable. Unfortunately, the paucity of exper
imental data on the tensile and compressive properties of cord-rubber
composites and the difficulty of u tili/.ing such properties, even if available,
in the analysis of a structure as complicated as a tire, prohibits the further
use of such a material property representation herein.
Multi-Ply Systems: Classical Lamination Theory
Now that the response of a single calendered ply of cord and rubber has
been characterized in both the specially and generally orthotropic configu-

10 30° 60° 90C


CORD ANGLE 9
FIGURE 3.13c Variation of transformed stiffness constants Q,h QI2, QI6 and QM with cord
angle 8 for 1000/2 polyester-rubber ply.
CORD REINFORCED RUBBER 147

On

30° 60° 90°


CORD ANGLE 9
FIGURE 3.14 Comparison of modulus E, with transformed stiffness Qn as a function of cord
angle 9 for 1000/2 polyester-rubber ply.

rations in terms of its various elastic constants (engineering, compliance


and stiffness constants), the response of multi-ply or laminated systems
used in tires can be investigated. A multi-ply system is composed of indi
vidual plies or layers at various cord angles bonded together in a pre
scribed stacking sequence. Each layer is assumed to be homogeneous but
is allowed to be anisotropic if required. The properties of each ply form
the basic building blocks which, together with the principles of mechanics
of materials, serve to define the material properties of the system. The sim
plest analysis of such multi-ply systems employs the assumptions of plane
stress and is known as classical lamination theory.
When two or more plies are bonded together the characterization prob
lem becomes essentially one of structural mechanics rather than material
properties. The relations connecting stresses and strains in multi-ply sys
tems are complicated compared to one-ply systems in that laminates will
generally twist and bend, as well as stretch, when subjected to a tensile
load.
This induced bending in multi-ply composites can be explained by con
sidering the off-axis tension test of two separate one-ply systems—with
one ply at a cord angle of +0°, the other at —0°. When loaded by a uni
form tensile stress applied at each end of the specimen (Fig. 3.15), an in-
plane shearing strain y, , occurs in accordance with the shear coupling
term A, defined by Eqs. 10 and 1 1. These deformation patterns are mirror
images of each other in each layer. When the two piles are bonded to
gether (±0), the oppositely directed in-plane shearing stresses in each layer
produce out-of-plane twisting of the laminate. Such twisting is easily ob
served in multi-ply cord-rubber specimens when axially extended as well
as in meridional sections removed from bias ply tires. Hirano and Aka-
saka [42] have made experimental measurements of this twisting behavior
in uniaxially loaded two-ply laminates over a range of cord angles (±6)
which are in good agreement with theoretical predictions.
The strains developed in the multi-ply system are therefore due to a
148 MECHANICS OF PNEUMATIC TIRES

LOADED

UNLOADED

(b)
FIGURE 3.15 Deformation patterns that occur in off-axis loading of: (a) individual plies at
+0 and —9 cord angles; (b) two-ply ±0 laminate.

combination of both bending and stretching of the laminated structure.


The stretching (or membrane) strains are related to derivatives of the dis
placements in-the-plane of the structure while the bending strains are re
lated to derivatives of the displacements out-of-plane of the structure. It is
these combined strains that should be used in the stress-strain or Hooke's
law for the multi-ply system.
However, it is convenient when analyzing multi-ply systems in plate or
shell-like configurations to work with stress resultants (rather than
stresses) and moment resultants (rather than moments) which are also
called stress couples. This is because the stresses are assumed to be linearly
distributed through the thickness of each layer in thin elastic laminated
plates or shells and the use of statically equivalent stress and moment re
sultants eliminates variations with respect to thickness giving a two-di
mensional theory. Thus, the stress resultants represent average but stati
cally equivalent values of the normal and shearing stresses integrated over
the thickness of the laminate. The stress resultants Nf N(, and N^ have di
mensions of force per unit length and are shown schematically acting on
an element of a two-ply bias tire, e.g., in Fig. 3.16a. Similarly, the moment
resultants AY..., M(, and AY.,., have physical dimensions of moment per unit
length (force) and are shown acting on an element of the same tire in Fig.
3.16b.
The meridional direction (denoted by <<>) and the circumferential direc
tion (denoted by £) are the principal directions associated with the toroidal
shape of a tire (see Figs. 3.16a and 3.16b). These directions are determined
solely by tire geometry and are not necessarily the principal material prop
erty directions. At this point it should be noted that three distinct sets of
coordinate systems have been used to characterize properties and/or the
geometry of relatively thin-walled cord-rubber systems: (a) the longitudi
nal 1 and transverse 2 directions used with unidirectional, specially ortho-
tropic plies, (b) the x and y directions used with off-axis, generally ortho
CORD REINFORCED RUBBER 149

WHEEL AXLE
-L z
FIGURE 3. 16a In-plane stress resultants N^, N( and ff^ acting on laminated tire structure
(and pertinent nomenclature).

tropic plies, and (c) the meridional <f> and circumferential £ directions
associated with multi-ply tire systems3. It so happens that for a radial tire
sidewall, the meridional <j> direction of the tire structure is coincident with
the longitudinal 1 direction of the sidewall cords.
When the combined (bending and membrane) strains are substituted
into the Hooke's law for a multi-ply system, the stress resultants N+, Ne,
and N^ and moment resultants Mv M(, M# are related to the middle sur
face strains ej, ej, Y* and curvature changes K+, K(, K^ in the following
form when the tire element depicted in Figs, 16a and 16b is analyzed for
stress at a point. (Many algebraic manipulations have been omitted herein
but the details can be found in Refs. 11-12):

(3.23)

Equations 3.23 will be referred to as the stress-strain relations, constitu


tive equations or Hooke's law for multi-ply systems as used in pneumatic
The Ihickoea direction utociaied with each of thae coordinate systems is denoted by "3", ":" and "{", ropectively.
ISO MECHANICS OF PNEUMATIC TIRES

FIGURE 3.l6b Moment resultants M+, Mt and M^ acting on laminated tire structure.

tires. The particular kinematical form of the middle surface strains


e£() and curvature changes (K+, K(, K^) depends on the geometry of the
structure being analyzed and the simplest situation is realized in the case
of flat plates. Such strain-displacement equations associated with particu
lar geometric configurations (e.g. toroidal shells) are not discussed herein
but may be found in standard shell-theory texts [43]. Note that there are
1 8 different "elastic coefficients" in the most general case needed to specify
material properties for any laminate (six each of [Av], [B;/] and [£>„]) re
gardless of its geometric configuration—in contrast to the simpler situation
that existed for a single ply which could be characterized in terms of four
independent engineering, compliance or stiffness constants.
The matrix equivalent of Eqs. 3.23 is

"id

/»|2 02 "2t> «J
*>66
(3.23a)
M. Blt BI2 B, D16
Mt B\i 822 B2
Z>,6

which often appears in the literature of composite material mechanics in


the following abbreviated notation:

N A B €°
I
1 (3.23b)
M B j D K
CORD REINFORCED RUBBER 151

In these relationships, the components of the [X,y] submatrix (with di


mensions of force per unit length) are associated with stretching of the
cord-rubber composite (extensional stiffnesses), those of the [Dv] sub-
matrix (with dimensions of length times force) are associated with bending
and twisting of the composite (bending stiffnesses), while those of the [ #„]
submatrix (with dimensions of force) couple the bending-stretching re
sponse of the material (coupling stiffnesses). In the analysis of engineering
laminates, Reissner and Stavsky [44] were the first investigators to predict
coupling between bending and stretching responses in the case of layered
plates in 1961; Ambartsumyan [45] actually noted this phenomenon ear
lier in layered shells, but his work did not appear in the English language
until 1964.
The values of the individual elements of the [Au], [B,,] and [£>„] sub-
matrices for the multi-ply system depend, in general, on the number of
plies in the tire, the thickness of each ply, the orientation of each ply, the
stacking sequence of the ply, and, of course, the elastic constants and vol
ume fractions of the cord and rubber components of the ply; values of the
individual elements of the submatrices can be calculated in terms of the
transformed reduced stiffness [Qv] as follows:

-**-.)- 2

where tk is the thickness of the kth ply, zk is the distance from the geometric
middle surface of the laminate to the centroid of the kth ply, and N is the
total number of plies in the laminate.
Typical values calculated for the 1 8 elastic coefficients for a variety of
tire constructions are given in Tables 3.6 and 3.7. These particular values
were obtained at the tire crown. Since cord angle, cord end count, ply
thickness, etc. generally vary with position in the tire, the elastic coeffi
cients will possess different values at different locations (crown, sidewall,
shoulder, etc.).
In many cases the stiffness matrix for the 18 elastic coefficients can be
simplified:
a. If for every ply at +6 another identical ply at —0 is present regardless of
the stacking sequence, the terms A 16 = AK = 0 and the laminate is said
to be balanced. Such a composite system behaves orthotropically with
respect to inplane stresses and strains. Balanced laminations are uti
lized in most bias, belted-bias and radial passenger tires.
b. If for every ply at +6 a given distance f above the geometric middle
surface, an identical ply at —6 is the same distance (i.e., — £) below the
middle surface, D,,, = D26 = 0 and the laminate behaves orthotropically
in bending. Most tires of bias construction feature this characteristic.
c. When an even or odd number of plies is used that are laid up sym-
etrically with respect to the geometric middle surface of the composite,
the elements of the [Bv] submatrix are zero. Such a system is said to be
uncoupled or symmetric; that is, the stresses (σ) depend only on middle
152 MECHANICS OF PNEUMATIC TIRES

TABLE 3.6 Elastic coefficientsfor bias, belted-bias and radial tire constructions
"Al B"
B | D*^

14.9 32.7 0. 0.052 0.126 -0.288


32.7 85.5 0. 0.126 0.322 -0.722
0. 0. 33.1 -0.288 -0.722 0.125
Riftt Tir* Ift2
0.052 0.126 -0.288 0.015 0.034 0.
0.126 0.322 -0.722 0.034 0.088 0.
. -0.288 -0.722 0.125 0. 0. 0.034
51.3 145 0. -0.492 -2.56 -2.27
145. 470. a -2.56 -11.6
-7.19
-7.19
0. 0. 146. -2.27 -2.57
-0.492 -2.56 -2.27 0.206 0.578 0.070
-2.56 -11.6 -7.19 0.578 1.87 0.331
_ -2.27 -7.19 -2.57 0.070 0.331 0.581
142 56.2 0. 14.5 -1.79 -0.454
56.2 721. 0. -1.79 -23.7 -5.64
0. 0. 57.2 -0.454 -5.64 -1.79
14.5 -1.79 -0.454 1.95 0.335 0.030
-1.79 -23.7 -5.64 0.335 4.25 0.372
-0.454 -5.64 -1.79 0.030 0.372 0.342

surface strains (e°) and the moments (A/) depend only on curvature
changes (K)—see Eqs. 3.23. Symmetric laminates are used in the man
ufacture of many "aerospace type" rigid composites as well as in ply
wood with 0° and 90° layups (so-called cross-ply laminates). Note that
in a four-ply bias tire with identical plies, the stacking sequence +6,
—0, —0, +6 ("paired ply" construction) is uncoupled while the se
quence +0, —6, +6, —a usually employed is coupled.
These various simplifications which result in the constitutive equations
governing the stress-strain behavior of multi-ply systems are discussed by
Ashton, et al. [11], pp. 45-49, and Jones [12], pp. 156-172.
It is apparent that the stress-strain behavior expressed by the Hooke's
law for a laminate, Eqs. 3.23, is very complex due to the presence of 18
elastic coefficients. However, for those configurations where bending-
stretching coupling vanishes ([Bv] = 0) or can otherwise be shown to be
unimportant, extensional and nexural deformation problems can be
treated independently and laminate material characterization is consid
erably simplified. Further when all of the coupling stillnesses vanish (A 16,
A26, D16, DM, [Bij] = 0) for a multi-ply system, the laminate is specially or-
thotropic (in analogy with the specially orthotropic ply). For this limiting
case, there are five elastic constant (E+, E(, G^, v^, v^-—four of which are
independent—for cord-rubber composites constituting the pneumatic tire
structure:
a. the meridional Young's modulus E+
CORD REINFORCED RUBBER 153

TABLE 3.7 Elastic coefficientsfor different radial tire constructions


"A 1 B"
1
B jD

129. 399. 0. 2.36 -8.76 -5.26


399. 2443. 0. -8.76 -53.7 -32.0
0. 0. 400. -5.26 -32.0 -8.77
Nylon Monofil Body/ 102
Steel Belt 2.36 -8.76 -5.26 0.393 0.756 0.231
-8.76 -53.7 -32.0 0.756 4.63 1.41
_ -5.26 -32.0 -8.77 0.231 1.41 0.759
'203. 399. 0. 6.40 -13.1 -5.26
399. 2440. 0. -13.1 -80.6 -32.0
0. 0. 399. -5.26 -32.0 -13.0
Rayon Body/ 102
Steel Belt 6.40 -13.1 -5.26 0.786 0.999 0.347
-13.1 -80.6 -32.0 0.999 6.11 2.11
-5.26 -32.0 -13.2 0.347 2.11 1.00 .
~I67. 130. 0. 14.8 -4.25 -1.57
130. 790. 0. -4.25 -26.2 -9.20
0. 0. 132. -1.57 -9.20 -4.30
Rayon Body/ 102 T
PVA Belt 14.8 -4.25 -1.57 2.11 0.752 0.105
-4.25 -26.2 -9.20 0.752 4.55 0.613
-1.57 -9.20 -4.30 0.105 0.613 0.766
171. 138. 0. 14.7 -4.52 -1.63
138. 825. 0. -4.52 -27.3 -9.60
0. 0. 138. -1.63 -9.60 -4.48
Rayon Body/ 102
Rayon Belt 14.7 -4.52 -1.63 2.14 0.798 0.108
-4.52 -27.3 -9.60 0.798 4.75 0.639
-1.63 -9.60 -4.48 0.108 0.639 0.797

b. the circumferential Young's modulus E(

c. the in-plane shear modulus

d. the Poisson's ratio

e. the Poisson's ratio

where reciprocity requires that

and h is the total thickness of all plies.


Such simplifying assumptions, among others, were employed a priori in
the analysis of cord-rubber laminates as developed classically by Gough
154 MECHANICS OF PNEUMATIC TIRES

[8j4, Akasaka [9] and Tangorra [46], and utilized by subsequent investiga
tors—e.g., Robecchi [47] and Akasaka and Hirano [48]. Angle-ply lami
nates (pairs of identical ±0 plies) are especially amenable to this type of
simplified analysis—i.e. multi-layer properties are directly expressible in
terms of cord angles as well as cord and rubber elastic constants and vol
ume fractions if the laminate is specially orthotropic. Thus with this ap
proach it can be shown that for a pair of otherwise identical ±0 plies

G* = Ecvc sin20 cos20 + G,(\ - vc)


_ Ecvc sin20 cos20 + 2Gr(l - vt)
** ~ Ecvc sin40 + 4G,( I - vc)
and
E((ff) = £>/2 - ff)

Additional pairs of plies at the same or different cord angles with the same
or different constitutent properties are easily accommodated by this ap
proach which will be referred to as modified netting analysis.
In the limit of inextensible cords (Ec —» oo), the Young's moduli and
Poisson's ratios given above reduce to expressions referred to herein as
classical netting analysis:
E, = 4Gr(l - vr)(cot"0 - cot20 + 1)
E( = 4G>(1 - v,)(tan40 - tan20 + 1)
v* = cot20
PS, = tan20.
The elastic constants calculated from classical lamination theory (using
the Halpin-Tsai equations for the single ply properties) for a four-ply ny
lon-rubber composite (+6, -0, -6, +8) are plotted in Figs. 3.17a to 3.17c
for cord angles from 0° to 90°. Again, the cord angle 6 is the angle be
tween the cord and the applied load in a uniaxial tension test of the com
posite. The single ply properties predicted from Eqs. 1 1 are also shown for
comparison and were previously plotted in Figs. 3.12a and 3.12b.
Note that the important effects of a bias angle in a multi-ply system
compared to a one-ply system are to significantly increase both the shear
modulus for all cord angles other than 0° and 90° and the Poisson's ratio
for cord angles between 50° and 85°. This latter phenomenon indicates
that a significant Poisson mismatch exists between the body and belt of a
radial tire.
The predictions of modified netting anaysis for extensional and shear
moduli differ very little from those of classical lamination theory for the
four-ply cord-rubber system. For systems other than angle-ply laminates,
netting analysis (modified or classical) should be used with caution.
4 Gough credits mow of the analytical developments reported in [8] to his colleague! H. V. Wamwnght and H. J. Law-
ton.
CORD REINFORCED RUBBER

4-PLY (Ej
UJ

(Oe) = l62,900 psi

0° 30° 60° 90°


CORD ANGLE 9

FIGURE 3.17a Variation of Young's moduli Et and Ef with cord angle 6 for four-ply and
one-ply 840/2 nylon-rubber systems, respectively.

Poisson's ratios well in excess of one-half are easily measured on angle-


ply cord-rubber laminates. Comparison of the various theoretical predic
tions (classical and modified netting analysis and classical lamination the
ory) with experimental measurements made on ±8 two-ply steel cord-rub
ber composites are shown in Fig. 3.18.
Direct measurement of Young's moduli on multi-ply cord-rubber uni-

36000 •
o 4-PLY
v, 30000-
° 24000 ••
CO
3 18000-
2 12000
£ 6000+
UJ
£ o 0° 30° 60° 90°
CORD ANGLE 9
FIGURE 3. 17b Variation ofshear moduli G^ and Gfy with cord angle 8forfour-ply and one-
ply 840/2 nylon-rubber systems, respectively.
IS6 MECHANICS OF PNEUMATIC TIRES

30° 60°
CORD ANGLE 9
:IGURE 3.17c Variation of Poisson's ratios v& and vyf with cord angle 8 for four-ply and
one-ply 840/2 nylon-rubber systems, respectively.

ixial test specimens gives values that are lower (in some cases, much
lower) than values predicted theoretically. This increased compliance in
excess of that predicted by classical lamination theory is mostly due to the
presence of interply stresses which promote a reduction in the apparent
stiffness of cord-rubber laminates. This effect is especially pronounced in
narrow test specimens.
Few published data exist comparing theoretically predicted with experi
mentally measured elastic constants for cord-rubber laminates; some of
that which does exist is suspect because of lack of attention to specimen
design and method of load application resulting in a nonhomogeneous
strain state in the gage section of the laminate.
In summary, the five elastic constants (Et, E(, G^, v^, v^ and/or the 18
elastic coefficients (A,h BtJ, Dv) may be calculated for any type of pneu-

20t
CLASSICAL
ANALYSIS
MODIFIED NETTING
ANAL., CLASSICAL
LAMINATION THEORY
EXPERIMENT

30° 60C
CORD ANGLE 9 i
FIGURE 3.18 Comparison between theoretically predicted and experimentally measured
values of Poisson's ratio v^ as a function of cord angle 8 for two-ply 7 X 5 X 0.010 in. steel
cord-rubber angle-ply laminates.
CORD REINFORCED RUBBER 157

matic tire structure with the use of the concepts presented herein regard
less of the number of plies, cord materials, cord angles, etc. However,
these predicted material properties should be verified experimentally
when possible before using them in subsequent analyses.
Multi-Ply Systems: Interlaminar Deformations
The preceding treatment of cord-rubber laminates, employing the as
sumptions of plane stress, cannot analytically accommodate the effects of
interlaminar (interply) stresses. While such stresses may be legitimately ig
nored in some problems of tire mechanics—such as in the calculation of
the cord load distribution due to inflation pressure in bias ply tires—they
are important with regard to certain failure phenomena occurring in serv
ice. It is the interply stresses that are mainly responsible for fatigue in
duced delamination (i.e., ply separations) which are prone to occur in lo
calized regions containing ply endings (viz., belt edges and turn-up ply
endings).
When a multi-ply cord-rubber system such as a tire is loaded, interply
normal (o,) and shearing (r,,,{, r(t) stresses are generally present, often at
appreciable magnitudes, in addition to the in-plane normal (o..,, o,) and
shearing (r^) stresses previously considered in the classical lamination the
ory. The interlaminar stresses produce corresponding strains e^, >0t and y(f.
These interply stresses and strains serve to reduce the overall stiffness of a
laminated structure—i.e. make it more flexible under a given load which
in turn lowers natural frequencies of vibration and static buckling loads.
At the present time there is little agreement on the best way to incorpo
rate interlaminar effects in multi-ply structures short of three-dimensional
theory of elasticity. One popular method of accommodating interply shear
in the stress-strain equations governing laminated systems is to proceed as
in the derivation for the Hooke's law of classical lamination theory (Eqs.
3.23), but with additional stress resultants Q.., and (A introduced which
represent average values of the interlaminar shearing stresses T+{ and r(l,
respectively, integrated over the thickness of the multi-ply system. (In
tegrals involving the interlaminar normal stress <r are not relevant here.)
The stress resultants Q+ and Q(, like Nf, N( and N^, have units of force per
unit length and are shown schematically acting on an element of a tire in
Fig. 3.19 (cf. Figs. 3.16a and 3.16b). The governing stress-strain equations
for the laminated structure are then derived in matrix notation as:

AH AI2 AH Blt Bt2 Blt> 0 0


A 12 A-n A2b BI2 B-J2 B-u, 0 0
AH A26 Aft B16 B26 Bf* 0 0
M. Bn B12 BI6 Dn DI2 D)t 0 0
- (3.24)
M( B{1 B22 B2t DI2 D22 D2b 0 0 \
BU B* Bft DI6 DK DM 0 0
Qt 000000 k\A« *,M«j *
-Q*. 000000 Jk,M«5 k\A»
158 MECHANICS OF PNEUMATIC TIRES

FIGURE 3. 19 Interply stress resultants q± and Q( acting on laminated tire structure (see Figs.
16a-16b).

where .-(„, A45, and A55 are three additional elastic coefficients (supple
menting the 18 previously established), and A-,, A., are shear correction fac
tors (analogous to that used in calculating shear deflection in beams).
Thus, the constitutive equations of classical lamination theory (Eqs. 3.23)
have been augmented by expressions relating the interply stress resultants
Q. and Q( with the corresponding middle surface interply strains y£f and

1*1
fi IrSJ
Values for the additional elastic coefficients A^ A45, and A55 of the lami
nate ultimately depend on the thickness shear moduli C»\, and Ci,, of the
individual plies and transform as second rank tensors [2]. However, mea
surements to determine the thickness shear moduli are difficult to perform
and no experimental data exist for cord-rubber composites, although to a
first approximation it may be assumed that Gfl = Gyz = Gxy.
The above summary of interply shear is based on the work of Whitney
[49]. Similar treatment of so-called shear-flexible structures have received
much attention in the literature of composite material mechanics in the
last decade. Ambartsumyan's classic text [50] develops in considerable de
tail the theory of shear-flexible plates; this subject is also analyzed by Ash-
ton and Whitney [51], and concisely reviewed by Bert [52]. The text of
Vinson and Chou [53] covers the derivation of the equations governing the
influence of interlaminar stresses on anisotropic plate and shell structural
response. This text, like most of the literature treating this subject, uses the
terminology "transverse shear" ("thickness shear" is also used) when re
ferring to the stress components T^ and T(C, rather than interlaminar or in
terply shear employed herein; this latter terminology is preferred for tires
since it is more descriptive of laminated structures and avoids confusion
with the transverse direction used to denote the direction perpendicular to
the cords in the plane of a ply.
For the purposes of this section, many of the interesting phenomena as
sociated with interply shear can be studied using a simplified approach to
analyze an angle-ply laminate subjected to uniaxial extension without re
CORD REINFORCED RUBBER 159

sorting to the mathematical complexities required for a more rigorous


elasticity solution. The analysis which follows is due to Kelsey [54], but is
similar in many respects to that of other investigators [42, 55-57].
Consider a two-ply ±0 cord-rubber laminate of width 2b and thickness t
between plies subjected to a longitudinal load as shown in Figs. 3.20a-
20b. Each ply has identical properties. In this model the interply deforma
tions are constant through the thickness t and occur only in the rubber be
tween the bottom surface of the cord in the upper layer and the top sur
face ofthe cord in the bottom layer. To a first approximation such a model
represents the behavior of the belt of a radial tire. Due to the applied load,
a uniform normal strain e, constant over the belt width and an interply
shearing strain y,,(y) varying over the belt width are produced. The other
strain components are not of direct interest for our purposes. (The other
component of interply shearing strain y,.. vanishes due to symmetry, the
in-plane shearing strain y,, varies across the belt width and linearly
through the thickness, the in-plane normal strain e, varies across the belt
width, and the interply normal strain e. is assumed to be negligibly small.)
By equilibrium or strain energy considerations, the interply shearing strain
y ., is calculated to be distributed over the width of the belt in the rubber
between the cord layers as follows:
- l/ism2ff) fit sink py
(3.25)
€, G, + ViEcvjw*9 cosh fib
where
JJ _j_ +
_ 12 2G, +
Equation 3.25 indicates that the interply shearing strain is maximum at
the free edge of the belt (y = ±b), decays exponentially to a small value a
short distance from the belt edge, vanishes along the belt centerline (y =
0), and is of opposite sign at each belt edge. These observations are quali
tatively supported by measurements made by Bo'hm [55] who inserted

LOAD

z =0

xy
(b) INTERPLY
RUBBER
(a) ±8 LAMINATE
FIGURE 3.20 Modelfor interply deformations (a) ±8 laminate subjected to longitudinal load,
(b) resulting strains on interply rubber.
160 MECHANICS OF PNEUMATIC TIRES

straight pins normal to the ply surface in a cord-rubber laminate model


belt system. The pins rotated under load to provide a visual observation of
the interply shearing deformation which occurs due to belt extension.
In the limit of inextensible cords (Ec →> oo), the magnitude of the inter-
ply shearing strain at each edge of the belt (y = ±b) is given by the rela
tively simple expression

which indicates that interply shear vanishes when two plies are oriented at
0 = ±54.7° (in addition to orientations of ±0° and ±90° using Eq. 25).
This cord angle (±54.7°) is that for which in-plane normal stress and
shearing strain are uncoupled—i.e., each off-axis ply responds in a spe
cially orthotropic fashion rather than in a generally orthotropic fashion.
Further, the reinforcing cords in the belt are unloaded along their axis
when oriented at ±54.7°- e., are neither in tension or compression when
the belt is extended [48].
Kelsey's approach can also be used to calculate the cord strain distribu
tion across the belt [54]; it has been extended by Turner [58] to account for
different cord and rubber properties in each ply, varying cord angles in the
plies, laminates of more than two plies, as well as for predicting the
Young's modulus and Poisson's ratio of laminates as influenced by shear
flexibility and belt width. This extended analysis, of necessity, includes the
transverse (or meridional) component of interply shearing strain y«
(which vanishes in balanced angle ply laminates) in addition to the cir
cumferential component •/„ considered in the simpler analysis.
It is relatively straight-forward to observe the maximum value of inter
ply shearing strain that occurs during bending or stretching of a multi-ply
cord-rubber composite by scribing a straight line on the edge of the speci
men and monitoring the rotation of the line under load. The magnitude of
the interply shearing strain is given by the tangent of the angle formed be
tween the initially unloaded line and the displaced loaded line—e.g., a 45°
line rotation, which is easily produced in some laminates for moderate belt
extensions (e, ~ 10%) corresponds to a relatively large interply shearing
strain (y,M ~ 100%). Such observations were first made by Böhrrv [55] in
1966 and later by Brakel [59] in 1974. Photographs of deformation pat
terns based on this approach are shown in Figs. 3.21a-21c for 1000/3 pol
yester-rubber laminates of two-ply (±60°) and four-ply constructions fea
turing different stacking sequences (normal +60°, -60°, +60°, -60° and
paired ply +60°, -60°, -60°, +60°) with each specimen extended approx
imately 10%.
Interply shearing strain measurements made at the free edge of two-ply
1000/2 polyester-rubber specimens based on data of Lou and Walter [60]
are shown in Fig. 3.22 for 10% specimen extension as a function of cord
angle (±0) along with the theoretical predictions of Kelsey [54], Eq. 3.25,
and Puppo and Evensen [57]. These experimental results are based on x-
ray measurements which provide the internal distribution of y ... across the
width of the specimen and are discussed in more detail in [60].
Further analyses of the interply stresses occurring during uniaxial ex
tension of laminated strips are reviewed by Jones [12], pp. 210-223 and
Grimes and Greimann [61] including discussion of the interply normal
stress a, (the so-called "peeling" stress). The sign of the peeling stress (ten
sile or compressive) is controlled by ply stacking sequence and it reaches
large values where a significant Poisson mismatch occurs between layered
CORD REINFORCED RUBBER 161

FIGURE 3.21 Edge view of 1000/3 polyester-rubber laminates subjected to 10% extension
illustrating interply shearing strain: (A) two-ply ±60° cord angles; (B) four-ply +60°, -60°,
+60°, -60° cord angles; (c) four-ply +60°, -60°, -60°, +60° cord angles.

systems (such as between the body and belt of radial tires). The sign of the
peeling stress also has a large effect on the ultimate and/or fatigue
strength of laminates [62].
The role of interply shear deformations in bias tires has been discussed
by Biderman, et al. [10], pp. 55-61, in which the flexural response of a
three layer strip is analyzed (rigid top and bottom layers separated by a
shear flexible core) showing, among other things, that interply shear be
comes of less importance in controlling bias tire performance as the num
ber of plies is increased.

KELSEY L54)
___ PUPPO a EVENSEN [571
• EXPERIMENTAL [601

20° 40°
CORD ANGLE 9
FIGURE 3.22 Comparison between theoretical predicted ( , Kelsey "; —, Puppo and
Evensen*4) and experimentally measured (9, Lou and Walter68) values of interply shearing
strain ytl as a function cfcord angle 8 for two-ply 1100/2 polyester-rubber laminates at 10%
specimen extension.
162 MECHANICS OF PNEUMATIC TIRES

Finally it is interesting to note that the three components of inter-


laminar stress in a tire correspond to the three basic modes of crack sur
face displacements in fracture mechanics for a separation oriented in the
$f plane (af-opening mode; T<f-forward shear mode; ^-parallel shear
mode).

3.1.4. APPLICATIONS OF CORD-RUBBER


COMPOSITE THEORY TO TIRES
Obstacle Enveloping Characteristics
One of the outstanding features of the bias tire construction compared
to the belted-bias or radial construction is its ability to envelope (or "swal
low") road obstacles and road irregularities such as stones and pavement
expansion joints. This phenomenon can be partially explained by treating
that portion of the tire in contact with the road surface (that is, the foot
print) as a laminated anisotropic beam subject to bending moments such
that so-called cylindrical bending occurs without the development of ap
preciable in-plane forces.
For obstacles such as pavement expansion joints which are aligned me-
ridionally in the tire footprint (that is, parallel to the wheel axle), a cir
cumferential moment is primarily induced locally in the cord-rubber ma
terial in the footprint. Thus, the pertinent terms from Eqs. 23 are (under
the assumptions of cylindrical bending):
N( - Avg + B»K( - 0, or c? - -
and

so that, upon substituting for e£ into the last expression above, we obtain
A/£ = (Dv - B^/A^Kf (3 .26)
Equations 3.26 is nothing more than the moment-curvature relation gov
erning the deflection of beams which is derived in elementary mechanics
of materials but which now includes the effect of anisotropy. The terms in
parentheses in Eq. 3.26 are equivalent to the beam flexural rigidity (usu
ally denoted by EI for isotropic materials, where E is the Young's modulus
of the beam material and I the area moment of inertia about the neutral
axis). The flexural rigidity is a measure of beam stiffness since it involves
both the material properties and cross-sectional geometry of the beam.
Thus, the circumferential flexural rigidity ( £7), for the cord-rubber lami
nates used in tires can be expressed as
D» - Bj/An. (3.27)
Note that for a homogeneous, isotropic material

Eh/(\ - i?)
CORD REINFORCED RUBBER 163

so that the familiar expression

is recovered (for a "wide" beam of depth h).


In matrix notation, the expression [Dv] — [BV](AV]~1[BV] denotes the so-
called reduced bending stiffness of a laminated anisotropic structure. The
use of the reduced bending stiffness to obtain approximate solutions to un-
symmetrically laminated plate deflection, buckling and vibration prob
lems is discussed by Ashton and Whitney [52], pp. 137-140.
Similarly, in the meridional direction
(EI)^Dn-B^/An. (3.28)
We now consider the problem of ranking tires of different construction
for obstacle enveloping characteristics under static loading conditions.
This method of predicting obstacle enveloping characteristics must be
viewed as being approximate because two important physical phenomena
have been neglected in the preceding analysis: (a) the influence of tire in
flation pressure and (b) the effect of differences in tensile and compressive
moduli in cord-rubber composites. Both factors are important in deter
mining the absolute values of cleat envelopment loads for tires but, even
when neglected, trends in such a response can be evaluated due to changes
in the internal structure of the tire.
Consider a four-ply bias tire with a 58° crown angle (used here as angle
between meridian and cord at crown rather than circumference and cord
as also frequently used), a 2 + 2 rayon/rayon belted-bias tire with crown
angles of 58° and 62° in the body and belt, respectively, and a 2 + 4
rayon/rayon radial tire with a 75° crown angle. The 18 elastic coefficients
evaluated at the crown for each tire construction are displayed in Table
3.6.
Comparison of the circumferential flexural rigidities of the tires given
by Eq. 3.27 with obstacle envelopment loads measured statically at 24 psi
inflation pressure on a 2 in. diameter semicircular steel cleat located me-
ridionally in the center of the footprint (Fig. 3.23) is given in Table 3.8.
The cleat is wider than the footprint of the tires and envelopment is con
sidered to occur when the tread rubber fore-and-aft of the cleat makes
contact with the flat surface to which the cleat is attached.
The low flexural rigidity of the bias tire in the tread region compared to
the two belted tires is evident both from the calculations and measure
ments. Only for the radial tire does the cleat enveloping load (1440 lb)
exceed the rated load for these tires at 24 psi inflation pressure (1380 Ib).
These large differences in the calculated rigidities can be easily demonstrated
qualitatively by building flat pads of cord and rubber which duplicate the
body and belt plies of these tires in the tread region and flexing the pads
with opposite sides of the pad gripped in each hand.
Other applications of Eqs. 3.27 and 3.28 to obstacle envelopment char
acteristics of tires are discussed by Walter, et al. [13].
Tread Wear Resistance
Belted tires, especially radials, have a significant advantage with respect
to rate of tread wear compared to non-belted bias tires. This large increase
in wear resistance in the belted constructions (with, importantly, no de
crease in traction properties) is larger than can be achieved by tread rub
ber changes alone; that is, wear improvements can be more easily effected
164 MECHANICS OF PNEUMATIC TIRES

-TIRE

2" DIA. CLEAT

FIGURE 3.23 Diagram for comparison of circumferentialflexural rigidity of tires of different


construction with meridional cleat envelopment loads (see Table 8).

by tire construction variations than by any practical choice of tread poly


mer and/or compound variations.
Wear of passenger tires occurs mainly when a vehicle is taken around a
curve during a cornering maneuver [63]. The tread center line distortion
which occurs in the footprint because of cornering is shown schematically
in Fig. 3.24. Slippage of the tread rubber relative to the road surface oc
curs mainly in the region at the rear of the footprint where the rubber re
turns to its unconstrained position. Belted tires experience less slippage
than non-belted tires in cornering. Even in straight-ahead driving (0° slip
angle) where little wear occurs, the belt appreciably reduces tread squirm
and produces a more uniform footprint pressure distribution in com
parison to non-belted constructions. Again, in a manner similar to that
used to study the obstacle enveloping characteristics of tires, this differ
ence in performance due to constructional variations can be explained by
treating the tire footprint as a laminated anisotropic beam subjected to a
meridional load (the cornering force) in the plane of the footprint.
Gough [8] recognized that a simple beam model could be employed to
relate tire construction features to tread wear rankings as early as 1952-53.
However, Gough did not use laminated anisotropic beam theory since this
DIRECTION
OF AIMED DIRECTION
MOTION

SLIP ANGLE

CORNERING
FORCE
TREAD
CENTERLINE

SLIP REGION

FIGURE 3.24 Tread center line distortion in tire footprint due to cornering.
CORD REINFORCED RUBBER 165

subject was virtually unknown twenty-five years ago. He did, however, in


clude shearing as well as the usual bending deformations in this beam cal
culations—i.e., for a load P centrally applied to a simply supported beam
of length L with elastic constants E and (/', the maximum deflection S is
given by

o„ m PU 2PL
48£7 %AG
where A is the cross-sectional area and I the area moment of inertia of the
beam. (Note that we are now considering in-plane loading of the tire foot
print in contrast to out-of-plane bending applicable to obstacle envelop
ment treated previously.) Recasting the above equation in the form of a
stiffness S (the so-called Gough stiffness), and substituting typical tire di
mensions for the quantities A, I and L (which Gough related to the relaxa
tion length of a tire), one obtains

where E( and G^ are circumferential and shear moduli of the cord-rubber


laminate in the tread region and C1, C2 are constants. (Gough used C1 =
3.85 in.-1 and C2 = 66 in.-1 which, in general, apply to the smaller Euro
pean size tires rather than the larger USA sizes). This type of analysis,
based on a mechanics-of-materials approach, has recently been extended
by Yamagishi and Takahashi [64] to embrace the effects of circumferential
tension (due to inflation pressure prestress) in the tread region of belted
tires. It should be kept in mind, however, that braking and driving forces
may be just as important (if not more important) than cornering forces in
influencing tread wear of tires other than passenger (e.g. aircraft, tractor,
earthmover, . . .); for such tire applications, the Gough model as originally
developed and used herein is inapplicable.
Depending on test conditions, passenger tires—especially belted con
structions—may wear at different rates at the shoulder and crown. It
should be appreciated that the beam model used in the anlaysis governs
crown rather than shoulder wear characteristics because of the influence
of belt edge properties not properly accounted for in the model. However,
we will use Gough's equation as originally derived (although it can be re
fined by employing lamination theory, accounting for inflation pressure,
and considering the sidewall as an elastic foundation which supports the
belt) because it seems to give all of the essentially correct trends regarding
the effect of tire construction features on tread wear. For example, Eq. 29
can be used to show that:
a. Radial tires with purely circumferential belts (90°) will wear poorly
because the shear modulus G^ of the cord-rubber composite is very
small even though the Young's modulus E( is maximized.
b. The optimum belt angle for radial tire wear resistance is in the
neighborhood of 70° to 75°, but this optimum angle depends on belt
cord modulus (to a large extent), belt skim stock modulus (to a lesser
extent) and tire dimensions.
c. The best wearing four-ply bias tires should have crown angles of
±45° and ±90°. (The 45° plies maximize G* while the ±90° plies
maximize E().
166 MECHANICS OF PNEUMATIC TIRES

d. Body cord and rubber moduli can have a significant effect on the
wear resistance of belted-bias tires but have little effect on the wear
of radial tires.
e. The relative wear ratings of a steel belted radial tire, a textile belted
radial tire, and a bias tire are, approximately, 200, 160, and 100, re
spectively [65].
Correlation of actual wear rates of tires of bias, belted-bias and radial
construction with Cough stiffness has been conducted by Daniels [66].
We now consider an example which demonstrates the application of Eq.
3.29 to the problem of predicting the wear resistance rankings of a radial
tire as influenced by a belt construction variable. Consider an HR78-15 2
+ 2 rayon/steel radial tire with varying belt angle. (The elastic coefficients
for this tire are given in Table 3.7 for a 68° crown angle.) In the spirit of
Gough's original work, the elastic constants E( and G^ appearing in Eq.
3.29, are determined directly from the elements of the [Av] matrix, that is,
E( = (AUA22 — A2n)/hAu and G^ = A^/h, where h is the total thickness of
the body and belt plies. Again, these expressions are strictly applicable
only when bending-stretching coupling is neglected ([Bv] = 6). A plot of
the calculated Gough stiffness S is shown in Fig. 3.25 for belt angles from
60° to 80°. Note from the curve that a maximum in tread wear resistance
occurs in the vicinity of 72-74° which is in agreement with experimental
results. This optimum angle changes with any cord-rubber parameter that
influences the circumferential modulus E( and the shear modulus G,^ of
the composite material used in the belt region, and with tire parameters
that influence the values of the constants C, and C2 appearing in Eq. 3.29.
In summary, the stiffness of a tire as influenced by laminate elastic con
stants is one of the principal factors influencing tread wear. The more rigid
the tire in the footprint, the greater the Gough stiffness S and hence, the

-12000

CO
:• 1 1 ooo
10000
en
CO
UJ
9000

H 8000
CO

5 7000
ID
8 6000

60° 64° 68° 72° 76° 80°


BELT ANGLE 9
FIGURE 3.25 Variation ofGough stiffness S with belt angle 8 for rayon/steel radial tire.
CORD REINFORCED RUBBER 167

better the tread wear resistance. Other applications of this concept—i.e.


ranking tires as a function of construction features for tread wear resis
tance—are given elsewhere [8, 13].
Vibrational Characteristics of Radial Tires
Pneumatic tires impose a variety of natural and forced vibrations upon
a vehicle which affect its ride characteristics. Low frequency forced vibra
tions of the tire (<20 Hz) are governed primarily by the spring rate behav
ior (slope of vertical load-deflection curve in operating region) of the in
flated, rolling tire and are adequately damped by the vehicle suspension
system. The higher frequency vibrations of the tire (>20 Hz), while of
lower intensity than the vibrations governed by the tire spring rate, are
usually lightly damped, but in some cases may be amplified because of lo
cal resonances occurring in the vehicle. These higher modes of vibration,
in theory, occur at an infinite number of discrete frequencies and are gov
erned by overall tire constructional and geometrical features. These tire
resonances may be road-induced due to, for example, pavement expansion
joints and other road irregularities or tire-induced due to, for example,
small force and moment variations occurring in the tire.
The purpose of this section is to assess the influence of belt material
property variations on the vibrational response of radial tires. We will
show that belt composition, while potentially variable over a wide range,
has surprisingly little influence on the natural resonant frequencies of
radial tires. This behavior is in marked contrast to the enveloping and
wear characteristics previously considered. (The material properties of the
belt do, however, influence vibrational amplitude.) To show the effect of
belt properties on frequency response, we note that the differential equa
tion of free flexural vibrations for a homogeneous, isotropic, cylindrical
shell of radius R which governs the out-of-plane displacement w of the
shell is given by [67]
d_ El
w—
a/2 mR" a? af if
where t is the time coordinate, £ is the circumferential space coordinate, m
is the mass per unit area, and EI is the flexural rigidity. Assuming a suit
able harmonic solution for this equation, the n resonant frequencies /„ are
given by
1

where (EI)( is the circumferential flexural rigidity of the belt denned by


Eq. 3.27. The fundamental mode of flexural vibration is given by n = 2,
since when n = 1 the belt moves as a rigid body.
Equation 3.30, as written, predicts frequencies much lower than those
measured on radial tires because it neglects, among other factors, the stiff
ening effects in the meridional and circumferential directions of the side-
walls, which act as an elastic foundation for the tread, and the tire infla
tion pressure which imposes a tensile pro-stress on the belt. These two
factors, which are independent of the composition of the belt, are domi
nant at low mode numbers (n = 1 to 6) in controlling the vibrational re
sponse of radial tires [68]. Thus, while the circumferential flexural rigidity
168 MECHANICS OF PNEUMATIC TIRES

TABLE 3.8 Comparison ofcircumferentialflexural rigidity with meridional cleat envelopment


loadfor different tire constructions (Fig. 3.23)
Circumferential Flexural Meridional Cleat
Rigidity, in-lh. Envelopment Load.
Tire Construction (EI){ - D^/A^ lb
8.25- IS Bias
Four-Ply Nylon 10 1210
G78-1S Belted-Bias
2 + 2 Rayon/Rayon 160 1320
20SR-1S Radial
2 + 4 Rayon/Rayon 350 1440

of the belt (I'-I\ can be made to vary by an order of magnitude (from


about 70 in · lb for a two-ply PVA belt to about 700 in · lb for a four-ply
fiberglass belt), this large variation in bending stiffness causes no change
in the frequency response of a radial tire at the lower modes of vibration.
This fact has been confirmed experimentally by Potts [69] using time-aver
age holographic techniques to display the resonant mode shapes of vibra
ting radial tires of varying belt composition. Potts [70] has further shown
that the frequencies of the lower modes of vibration of "beltless" radial
tire bodies are almost identical to those of their belted counterparts. Table
3.9 lists the first six modes of vibration measured at 24 psi inflation pres
sure for two otherwise identical FR70-14 tires featuring a two-ply steel
belt and a four-ply rayon belt. At the higher modes of vibration (n > 6),
the circumferential flexural rigidity becomes important compared to side-
wall stiffness and belt prestress, and radial tires with different belt con
struction features will vibrate at appreciably different frequencies. These
frequencies (>200 Hz) are in the airborne tread noise spectrum.
The important observation from this kind of study is that belt construc
tion features (both cord angle and material properties) have little effect on
the frequency response characteristics of a radial tire in the region of inter
est to automotive engineers (with other factors, such as tire size, remaining
unchanged).
Stress Analysis
The simplest stress analysis problem to consider from a theoretical ap
proach deals with the calculation of the shape taken by and the stresses
developed in an inflated but otherwise unloaded tire. The first rigorous SO

TABLE 3.9 Resonantfrequencies ofFR70-14 radial tires


Frequency, Hz
Two-Ply Rayon Body, Two-Ply Rayon Body
Mode Number Two-Ply Steel Belt Four-Ply Rayon Belt
1" 65 66
2 78 . 78
3 % 99
4 118 118
5 135 140
6 157 162
' Rigid body motion
CORD REINFORCED RUBBER 169

lution to this problem for the bias construction was obtained by Purdy in
1928 [71]. The importance of this shape equation is due to the fact that
once the geometry of the inflated tire is accurately known, the stresses de
veloped in it can be calculated. We now show how Purdy's equations,
though seemingly independent of cord-rubber composite material proper
ties, can be extended to embrace any type of tire construction if certain
elastic constants are introduced into the anlaysis.
If it is assumed that the inflation pressure p normal to the inside surface
of the tire is balanced only by the internal reactions N+ and Nf at the
middle surface of the laminated tire structure (see Fig. 3.16a), equilibrium
exists, so that summing forces normal to the surface leads to

.
N. re N.
where r^, and r( are the principal radii of curvature for the doubly curved
toroidal surface. Implicit in obtaining this equation is the assumption that
the in-plane shearing stress resultant N^ vanishes by symmetry, which is
approximately true for balanced laminates (i.e., most bias, belted-bias and
radial ply tires). However, whenever any appreciable unbalance occurs
due to construction (AI6 ^ 0, A26 ^ 0), in-plane shear is most certainly
present in the plies of the inflated tire and /V,x must be included in the
equations of equilibrium—i.e., the structural response of the tire to an in
flation pressure loading is non-axisy mmetric due to the material properties
of the plies. This somewhat unexpected situation is analogous to that
which occurs in uniaxial extension of a single off-axis (generally ortho-
tropic) ply of cord and rubber—i.e., normal stress-shearing strain coupling
is present. (Lekhnitskii's monograph [2] contains many examples of the
unusual deformation patterns that occur in anisotropic elastic bodies un
der simple loadings.)
Another equation of equilibrium can be obtained by summing forces
along the axis of revolution (wheel axle) to obtain:

(3.32)

where r is the radial coordinate measured from the axis of revolution to an


arbitrary point on the tire meridian and r, is the radial distance from the
axis of revolution to the widest part of the tire (see Fig. 3.16a).
The principal radii of curvature for a surface of revolution such as a to-
roid in rectangular Cartesian coordinates expressed in differential form
are given by

and

where z is the axial coordinate (which coincides with the wheel axle), and
z′ and z" denote the first and second derivatives of z with respect to r, re
spectively. Eliminating N+ from the right hand side of Eq. 3.31 (using Eq.
170 MECHANICS OF PNEUMATIC TIRES

3.32), substituting for the principal radii of curvature, integrating once,


and recognizing, with tan <j> = z′, that
sin <t> = z'[l + (*OT"2 (3.33)
leads to the expression,

(3.34)

where rc is the radial coordinate of the crown. Equation 3.34, in con


junction with Eq. 3.33, defines the inflated cross-sectional geometry of any
tire of balanced ply construction in rectangular Cartesian coordinates, z =
z(r). (Equation 3.34 is a generalization of an equation first derived by
Purdy and given as Eq. 3.16 in his monograph [71].) However, in order to
integrate Eq. 3.34, the principal stress ratio Nt/N^ must be prescribed.
In terms of balanced cord-rubber composite material properties (A „, =
AM, = 0), the principal stress ratio, neglecting bending and twisting (K. =
K( = K« = 0), is given by (see Eqs. 3.23)
Nt = t2 22
N. ( .....M
where use has been made of the following relations:

If the Poisson's ratios v^ and v^ and the middle surface strain ratio
are known for the cord-rubber laminated structure as a function of posi
tion r from the bead to the crown of the tire, the expression

exp [*L*L
J N. r
which appears in Eq. 3.34 can be integrated using Eq. 3.35. For belted
tires (belted-bias and radial), this integration is numerically complicated.
However, Robecchi [47], using similar procedures, has determined the in
flated cross-sectional profile and cord loads developed in a 165SR-13
radial tire featuring two rayon body plies and four rayon tread plies at 26
psi inflation pressure.
For bias tires, Eq. 3.35 can be considerably simplified. In the bias con
struction, the biaxial strains e", and e° resulting from inflation pressure (or
for other loads possessing rotational symmetry), are not independent of
one another under the assumption of pin-jointedness (pantographing).
Rather, the state of stress is such thai

and Eq. 3.35 reduces to

TF1-'* (3-36)
CORD REINFORCED RUBBER 171

Further, resolution of forces in the circumferential and meridional direc


tions at any point in the middle surface of the tire leads to expressions that
can be recast as

^ = tan20 (3.37)
"4
where 6 is the local cord angle with a meridian.
Comparison of Eqs. 3.36 with 3.37 indicates that
v& = tan20
which was previously developed as the expression for one of the Poisson's
ratios of angle-ply laminates in the limit of inextensible cords.
The above interrelationships form the basis of the classical netting anal
ysis of filamentary composite materials. Thus, in the work of Purdy [71]
and others, the approximation is made that the principal stress ratio N(/N^,
is given by tan20. The success of the netting analysis approximation in pre
dicting the inflated contour and cord loads in bias tires and in the sidewall
region of radial tires is due to the fact that Eqs. 3.36 and 3.37 produce ap
proximately the same results for cord angles in the range of commercial
interest (±0°, to ±65°) as is made apparent by inspection of Fig. 3.18.
Purdy [71] discusses in detail methods of integrating Eq. 3.34 for the
special case of bias tires where netting analysis is applicable. Much re
search remains to be done in this area for belted-bias and radial tire con
structions for which netting analysis concepts are generally inapplicable.
Ply Steer Behavior
A loaded rolling tire travelling restrained about its vertical (or yaw) axis
at zero camber and slip angle generally produces a net lateral force per
pendicular to the wheel plane of the tire.5 This lateral force is composed of
two components: conicity (or pseudo-camber) which does not change di
rection when the direction of rotation of the tire is reversed and ply steer
(or pseudo-slip) which changes direction when the direction of rotation of
the tire is reversed [72]. Conicity is caused by variations in manufacturing
tolerances from side-to-side of the tire and is not discussed further; ply
steer is associated with the inherent construction features of the tire and is
related to the intensity of bending-stretching coupling of the cord-rubber
composite [73]—i.e., the magnitude of certain elements of the Bu coupling
stiffness in the laminate stress-strain relations (Eqs. 23).
Bending-stretching coupling tends to cause a tire on a vehicle in service
to operate at a slip angle with zero ply steer. This in turn produces so-
called "dog tracking" on the vehicle (rear tires do not follow or track in
the path of the front tires) leading to, e.g., headlight misalignment. Mani
festations of such a slip angle are easily observed under laboratory condi
tions by deflecting a tire free to pivot about its yaw axis against a lubri
cated plate. The resulting angular distortion may be as large as 0.5° in
radial passenger tires but is usually less than 0. 1 ° in bias tires. If the yaw
axis of the tire is restrained from pivoting during rolling, as in most labo
ratory road wheel tests, the resulting lateral force is called residual corner
ing force and the resulting restoring moment is called residual aligning
torque.
While the ply steer force is generally low in bias tires, it can be made
5 A vertically directed restoring moment vector is also produced but is not considered herein.
172 MECHANICS OF PNEUMATIC TIRES

negligibly small by using paired ply construction (+6, —0, —6, +ff) for
which all elements of the B,j matrix vanish.
It is usually not practical to build belted tires for which the coupling
stiffness vanishes because of body-belt interactions. Pottinger [73] has
shown, however, that ply steer is predominately due to the presence of the
Bu term in the bending-stretching coupling matrix and secondarily to B166.
Thus, "belt packages" that minimize these two elements (especially B26)
generate low ply steer forces. This is demonstrated in Table 3.10 com
paring ply steer with the BfJ belt matrix for otherwise identical 2 + 4 poly
ester/rayon radial tires featuring three different rayon belt packages
placed on conventional 0° polyester body plies. The four-ply belt stacking
sequence "+6, +0, -0, -ff" produces the largest value of ply steer (96 lb.),
the sequence "+9, —6, +6, —0" produces an intermediate value (57 lb.),
and the sequence "+0, ^0, -0, +ff" produces the smallest value (19 lb.)—
all of which are consistent with the calculated magnitude of B,,,. (It should
be noted that these ply steer forces were measured on a 33 in. diameter
roadwheel and as roadwheel diameter increases the force measured de
creases appreciably; in any case, for these three constructions there is a lin
ear relation between ply steer and B26.)
It is well known that ply steer is larger in radial tires than in bias tires; it
is larger in radial tires with two steel belts than with four rayon belts.
These observations are in agreement with the values of the B26 terms of
Tables 3.6 and 3.7, respectively. Similarly, ply steer increases in steel bel
ted radials as the distance between belt plies (i.e., interply rubber) in
creases. This is in accordance with calculations for B26.
There are a variety of laminate configurations that have been proposed
to eliminate ply steer in tires, many of which have been patented espe
cially for radial tires [74-76]. fundamentally, ply steer is believed to arise
from a net inbalance in interply shearing action that can be generated dur
ing passage of the elements of the structural tire composite through the
footprint. Thus, even though interply shear is produced in paired ply con
structions, its effects are cancelled through the laminate thickness com
pared to normal four-ply constructions (cf. Figs. 3.21b-3.21c). Such inter
ply shear is in turn connected with the magnitude of the in-plane normal
stress-shearing strain coupling phenomenon generally present in each ply.
Calculations relating shearing action in the tire components during pas
sage through the footprint to ply steer related phenomena were conducted
by Longmore and Goodall [77] using equilibrium equations of elasticity
with modified netting analysis to represent cord-rubber material proper
ties. Ply steer, of course, vanishes in tires of cordless construction regard
less of whether they are pneumatic or solid.

3.1.5. LIMITATIONS OF THEORY—CONCLUDING


REMARKS
The equations presented in this section are theoretical and consequently
are somewhat limited because of the following simplifying assumptions:
a. Strains in the cord-rubber composite were assumed to be small and
linearly related to stress yet, under tire operating conditions, textile
Became of difierent labeling of meridional and circumferential directions. Potlinger'l [73] BK, a the same as our
and similarly for the other elemenu of the A<j, H,, and Du submatrices.
CORD REINFORCED RUBBER 173

TABLE 3.10 Ply steer behavior ofHR78-I5 2 + 4 Polyester/Rayon radial tires


B,, Belt Coupling
Bell Stacking Sequence Matrix (lb) Ply Steer Ob)*
+68°, +68°, -68°, -68° f 0 0 -332 T
0 0 -1415 %
L-332 -1415 0 J
+68°, -68°, +68°, -68° T 0 0 -166"!
I 0 0 -707 57
L-166 -707 0 J
+68°, -68°, -68°, +68° r o 001
000 19
0 00
" Measured on 33 in. diameter roadwheel

cord strains of several percent are developed at some locations in the


tire with even larger, perhaps nonlinear, strains being generated in
the rubber between the cords.
b. A condition of plane stress was used in the majority of equations re
lating stress and strain when, in fact, shearing deformations through
the thickness of the cord-rubber composite could be appreciable,
which will cause deviations from the plane stress state—i.e., inter-
laminar normal and shearing strains are present in cord-rubber com
posites that cannot be accommodated within the assumptions of a
plane stress (classical lamination) theory.
c. It was assumed that the compressive moduli for cord-rubber com
posites were the same as the tensile moduli when, in fact, the behav
ior of such systems generally are different in tension from in com
pression.
d. The rate of deformation and temperature effects associated with the
stress-strain behavior of viscoelastic materials were neglected hi both
cord and rubber.
e. A perfect adhesive bond was assumed to exist at the interface be
tween the rubber matrix and the reinforcing cords. If this bond rup
tures, tire performance is impaired and some of the cord-rubber
composite elastic constants can be considerably different from those
of the well adhered system.
f. All plies in the laminated tire structure were assumed to be in planes
parallel to one another for purposes of calculating elastic constants
although there are some regions in the tire where this condition is
not realized, viz. at the tum-up ply endings approaching the bead
and at the belt edges if insert rubber stocks are employed.
Thus, our present state of knowledge of the mechanics of laminated
structures is not sufficiently developed to allow reliable prediction of all
aspects of tire performance as influenced by the stiffness and strength
properties of cord-rubber composites. However, even with the above as
sumptions, certain trends in stiffness-dominated tire performance parame
ters may be predicted. This has been demonstrated through a discussion of
a number of tire mechanics problems including obstacle enveloping, tread
wear, vibration, stress analysis, and ply steer. The mathematical models
used to represent such tire behavior were sufficiently complicated to repre
sent reality but sufficiently simple to effect analysis. Even though emphasis
174 MECHANICS OF PNEUMATIC TIRES

was placed on the tire tread region for the applications discussed, other
areas of geometric and material complexity can be analyzed in a similar
manner. Finally, areas of tire research suggested by modern composite
materials mechanics can now be investigated on a more rational basis than
in the past.

3.1.6. ACKNOWLEDGEMENTS
The author of Section 3.1 (J. D. Walter) acknowledges with pleasure the
many stimulating discussions held on the subject of composite material
mechanics with his colleagues at Firestone, particularly H. P. Patel, J. L.
Turner and J. L. Ford, and with his associates in academia, particularly S.
K. Kelsey of the University of Notre Dame and C. W. Bert of the Univer
sity of Oklahoma. Additionally, many researchers in the USA and else
where, too numerous to name individually, have provided many refer
ences that were beneficial in the writing of this section. Finally, the author
wishes to thank the Management of The Firestone Tire & Rubber Co. for
creating the atmosphere which allows this work to nourish and for their
permission to publish it.

3.1.7. PRINCIPAL NOTATION


[A,,\ Extensional stiffnesses for a laminate
[/?,,) Coupling stiffnesses for a laminate
[D,] Bending stiffnesses for a laminate
£„ E, Young's moduli of cord and rubber, respectively
E,, E2 Young's moduli of specially orthotropic ply in longitudinal
and transverse directions, respectively
£„ Ey Young's moduli of generally orthotropic ply in x and y di
rections, respectively
£,,„ /•„', Young's moduli of laminated tire structure in meridional
and circumferential directions, respectively
Shear moduli of cord and rubber, respectively
Shear modulus (in-plane) of specially orthotropic ply
Shear modulus of generally orthotropic ply in xy plane
Shear modulus (in-plane) of laminated tire structure
Moment resultants
Stress resultants (in-plane)
Stress resultants (interply)
Stiffness matrix
Transformed stiffness matrix
Compliance matrix
Transformed compliance matrix
Cough stiffness (Eq. 3.29)
Cord volume fraction
Normal stresses in cord and rubber, respectively
Normal (a) and shearing (r) components of in-plane stress
for specially orthotropic ply
CORD REINFORCED RUBBER 175

o,, Oy, TX> Normal (a) and shearing (T) components of in-plane stress
for generally orthotropic ply
o» f*n T,, Normal (a) and shearing (T) components of interply stress
in angle-ply laminate
as> T«> T« Normal (a) and shearing (T) components of interply stress
in laminated tire structure
€„ e. Normal strains in cord and rubber, respectively
«i» ^2. Yi2 Normal (c) and shearing (y) components of in-plane strain
for specially orthotropic ply
«« «r» y,y Normal (e) and shearing (y) components of in-plane strain
for generally orthotropic ply
«« Y*« Yr, Normal (e) and shearing (y) components of interply strain
in angle-ply laminate
€f Y*f Y« Normal (e) and shearing (y) components of interply strain
in laminated tire structure
!>„ v, Poisson's ratios of cord and rubber, respectively
v,i, Vn Poisson's ratios of specially orthotropic ply
v,y, Vy, Poisson's ratios of generally orthotropic ply
»"«<> "tt Poisson's ratios of laminated tire structure
0 Angle between cord and x direction; angle between cord
and meridional direction in tire

References
[1] Bell, J. I ., The Experimental Foundations of Solid Mechanics, Encyclopedia of Physics,
Mechanics of Solids I, Vol. VIa/1, C. Truesdell. Ed., p. 17 (Springer, New York,
1973).
[2] Lekhnitskii, S. G., Theory of Elasticity of an Anisotropic Elastic Body (Holden-Day,
San Francisco, 1963). (Original publication in Russian, Government Publishing
House for Technical-Theoretical Works, Moscow, 1950).
[3] Hearmon, R. F. S., An Introduction to Applied Anisotropic Elasticity (Oxford Univer
sity Press, London, 1961).
[4] Martin, F., Theoretische Untersuchungen zur Frage des Spannungszustandes im Luf-
treifen bei Abplattung, Jahrbuch der Deutschen Luftfahrtforschung, Teil I, 470-4%
(1939).
[5] Clark, S. K., A Review of Cord-Rubber Elastic Characteristics, Rubber Chemistry and
Technology 37, 1365-1390 (1964).
[6] Clark, S. K., The Plane Elastic Characteristics of Cord-Rubber Laminates, Textile Re
search Journal 33, 295-313 (1963).
[7] Clark, S. K., Internal Characteristics of Orthotropic Laminates, Textile Research Jour
nal 33, 935-953 (1963).
[8] Gough, V. E., Stiffness of Cord and Rubber Constructions, Rubber Chemistry and
Technology 41, 988-1021 (1968); Kautschuk und Gummi Kunststoffe 20, 469-481
(1967).
[9] Akasaka, T., Various Reports and/or Bulletins of the Faculty of Science and Engineer
ing (Chuo University, Tokyo, 1959-1964).
[10] Biderman, V. L., et al., Automobile Tires, Construction, Design, Testing and Usage,
1969 (NASA Technical Translation F 12, 382). (Original publication in Russian,
State Scientific and Technical Press for Chemical Literature, Moscow, 1963).
[11] Ashton, J. E., Halpin, J. C. and Petit, P. H , Primer on Composite Materials: Analysis
(Technomic, Stamford, Connecticut, 1969).
[12] Jones, R. M., Mechanics of Composite Materials (McGraw-Hill, New York, 1975).
13] Walter, J. I) , Avgeropoulos, G. N., Janssen, M. L. and Potts, G. R., Advances in Tire
Composite Theory, Tire Science and Technology 1, 210-250 (1973).
[14] Treloar, L. R. G., The Physics of Rubber Elasticity, 3rd Ed., p. 87 (Oxford University
Press, London, 1975).
[15] Ferry, J. D., Viscoelastic Properties of Polymers, 2nd Ed. (Wiley, New York, 1970).
16] Kraus, G., Reinforcement of Elastomers by Carbon Black, Advances in Polymer Sci
ence 8, 155-237 (1971).
176 MECHANICS OF PNEUMATIC TIRES

[17] Nielsen, L. E., Mechanical Properties of Polymers and Composites, Vol. 2 (Marcel Dek-
ker, New York, 1974).
[18) Timoshenko, S. P., History of Strength of Materials, p. 216 (McGraw-Hill, New York,
1953).
[19] Novak, R. C. and Bert, C. W . Theoretical and Experimental Bases for More Precise
Elastic Properties of Epoxy, Journal of Composite Materials 2, 506-508 (1968).
[20] Massonnet, Ch., Two-Dimensional Problems of Elasticity, Handbook of Engineering
Mechanics, W. Fliigge, Ed., pp. 37-1 to 37-3 (McGraw-Hill, New York, 1962).
[21] Ward, I. M., Mechanical Properties of Solid Polymers, pp. 226-229 (Wiley-lnterscience,
New York, 1971).
[22] Brewer, H. K., Stresses and Deformations in Multi-Ply Aircraft Tires Subject to Infla
tion Pressure Loading, 1970 (AFFDL-TR-70-62).
[23] Whitney, J. M., Elastic Moduli of Unidirectional Composites With Anisotropic Fila
ments, Journal of Composite Materials 1, 188-193 (1967).
[24] Backer, S., Fibrous Materials, in Mechanical Behavior of Materials, F. A. McClintock
and A. S. Argon, Eds., pp. 675-706 (Addison-Wesley, Reading, Massachusetts,
1966).
[25] Wood, J. O. and Redmond, G. B., Tyre-Cord Behaviour Under Compressive Stresses,
Journal of the Textile Institute 56, T191-T204 (1965).
[26] Hearle, J. W. S., Grosberg, P. and Backer, S., Structural Mechanics of Fibers, Yarns,
and Fabrics, Vol. 1 (Wiley-lnterscience, New York, 1969).
[27] Takeyama, T. and Matsui, J., Recent Developments with Tire Cords and Cord-to-Rub-
ber Bonding, Rubber Chemistry and Technology 42, 159-256 (1969).
[28] Draves, C. Z., Lee, Z. S. and Skolnik, L., Survey of Cord Candidates for Radial Tire
Belts, Rubber World 164 (4), 41-47 (1971).
[29] Lehmicke, D. J., Firestone Tire & Rubber Co. Research Report (1976) (unpublished).
[30] Sims, D. F. and I lulpin, J. C., Methods for Determining the Elastic and Viscoelastic Re
sponse of Composite Materials, in Composite Materials: Testing and Design (Third
Conference), ASTM Special Technical Publication 546, pp. 46-66 (ASTM, Phila
delphia, 1974).
[31] Daniels, B. K., Steel Ribbon Belt Reinforcement Mechanics, Tire Science and Tech
nology 5, (1977).
[32] Hashin, Z., Viscoelastic Fiber Reinforced Materials, AIAA Journal 4, 1411-1417
(1966).
[33] Hashin, Z., Complex Moduli of Viscoelastic Composites-II. Fiber Reinforced Materi
als, International Journal of Solids and Structures 6, 797-807 (1970).
[34] Halpin, J. C. and Kardos, J. L., The Halpin-Tsai Equations: A Review, Polymer Engi
neering and Science 16, 344-352 (1976).
[35] Chamis, C. C. and Sendeckyj, G. P., Critique on Theories Predicting Thermoelastic
Properties of Fibrous Composites, Journal of Composite Materials 2, 232-358
(1968).
[36] Bert, C. W., Experimental Characterization of Composites, in Composite Materials,
Structural Design and Analysis, Vol. 8, Part II, C. C. Chamis, Ed., pp. 73-133 (Aca
demic, New York, 1974).
[37] Calcote, L. R., The Analysis of Laminated Composite Structures (Van Nostrand Rein-
hold, New York, 1969).
[38] Turner, J. L., Firestone Tire & Rubber Co. Research Report (1977) (unpublished).
[39] Tarnopol'skii, Yu. M., Roze, A. V. and Kintsis, T. Ya., The Bending of Clamped Beams
Made of Materials With Low Resistance to Shear, Polymer Mechanics 3, 486-491
(1967).
[40] Patel, H. P., Turner, J. L. and Walter, J. D., Radial Tire Cord-Rubber Composites,
Rubber Chemistry and Technology 49, 1095-1110 (1976).
[41] Bert, C. W., Models for Fibrous Composites With Different Properties in Tension and
Compression, Paper presented at 15th Midwestern Mechanics Conference, Univer
sity of Illinois at Chicago Circle, 23-25 March 1977; to be published in Journal of
Engineering Materials and Technology.
[42] Hirano, M. and Akasaka, T., Coupled Deformation of an Asymmetrically Laminated
Plate, Fukugo Zairyo (Composite Materials and Structures) 2, 6-11 (1973).
[431 Kraus, H., Thin Elastic Shells (Wiley, New York, 1967).
[44] Reissner, E. and Stavsky, Y., Bending and Stretching of Certain Types of Hetero
geneous Aeolotropic Elastic Plates, Journal of Applied Mechanics 28, 402-408
(1961).
[45] Ambartsumyan, S. A., Theory of Anisotropic Shells, 1964, (NASA Technical Trans
lation F- 1 18). (Original publication in Russian, State Publishing House for Physical
and Mathematical Literature, Moscow, 1961).
CORD REINFORCED RUBBER 177

[46] Tangorra, (>., Simplified Calculations for Multi-Ply Cord-Rubber Sheets as a Combi
nation of Orthotropic Laminates (Russian), Proceedings International Rubber Con
ference (Moscow, 1969), pp. 459-467 (Khimiya, Moscow, 1971).
[47] Robecchi, E., Mechanics of the Pneumatic Tire, Pan II, The Laminar Model under In
flation and in Rotation, Tire Science and Technology 1, 382-438 (1973).
[48] Akasaka, T. and Hirano, M., Approximate Elastic Constants of Fiber Reinforced Rub
ber Sheet and It's Composite Laminate, Fukugo Zairyo (Composite Materials and
Structures) 1, 70-76 (1972).
[49] Whitney, J. M., Stress Analysis of Thick Laminated Composite and Sandwich Plates,
Journal of Composite Materials 6, 426-440 ( 1 972).
[50] Ambartsumyan, S. A., Theory of Anisotropic Plates (Technomic, Stamford, Con
necticut, 1970). (Original Publication in Russian, Nauka, Moscow, 1967).
[51] Ashton, J. E. and Whitney, J. M., Theory of Laminated Plates, pp. 141-153 (Tech
nomic, Stamford, Connecticut, 1970).
[52] Bert, C. W., Analysis of Plates, pp. 188-192; Analysis of Shells, pp. 240-242, in Com
posite Materials, Structural Design and Analysis, Vol. 7, Part I, C. C. Chamis, Ed.
(Academic, New York, 1975).
[53] Vinson, J. R. and Chou, T. W., Composite Materials and Their Use in Structures
(Wiley, New York, 1975).
[54] Kelsey, S., Private Communication to Firestone Tire & Rubber Co. (1975).
[55] Böhm, F., Mechanics of the Belted Tire, University of Michigan ORA Translation No.
5, (1967); Ingeniur-Archiv 35, 82-101 (1966).
[56] Pipes, R. B. and Pagano, N. J., Interlaminar Stresses in Composite Laminates Under
Uniform Axial Extension, Journal of Composite Materials 4, 538-548 (1970).
[57] Puppo, A. H. and Evensen, H. A., Interlaminar Shear in Laminated Composites Under
Generalized Plane Stress, Journal of Composite Materials 4, 204-220 (1970).
[58] Turner, J. L., Firestone Tire & Rubber Co. Research Report (1977) (unpublished).
[59] Brakel, H., Model Studies on Tyre Materials, Kautschuk und Gummi Kunststoffe 29,
132-136 (1976).
[60] Lou, A. Y. C. and Walter, J. I) , Interlaminar Shear Strain Measurements in Cord-Rub
ber Composites. To be presented at SESA meeting, Wichita, Kansas May 1978.
[61] Grimes, G. C. and Greimann, L. F., Analysis of Discontinuities, Edge Effects and
Joints, in Composite Materials, Structural Design and Analysis, Vol. 8, Part II, C. C.
Chamis, Ed., pp. 158-170 (Academic, New York, 1975).
[62] Pagano, N. J. and Pipes, R. B., The Influence of Stacking Sequence on Laminate
Strength, Journal of Composite Materials 5, 50-57 (1971); (AFML-TR-71-8).
[63] Gough, V. E. and Shearer, G. R., Front Suspension and Tyre Wear, The Institution of
Mechanical Engineers, Proceedings of the Automobile Division, pp. 171-216 (1955-
56).
[64] Yamagishi, K. and Takahashi, S., Radial Tire Having High Modulus Breakers, United
States Patent 3,821,977 (1974); British Patent 1,310,316 (1973).
[65] Gough, V. E., Nondestructive Estimation of Resistance of Tire Construction to Treat
Wear, SAE Paper No. 667A, (1973).
[66] Daniels, B. K., A Calibration Curve for Treadlife as a Function of Gough Stiffness,
SAE Paper No. 770329, (1977).
[67] Young, I )., Vibrations of Continuous Systems, Handbook of Engineering Mechanics,
W. Fliigge, Ed., p. 61-19 (McGraw-Hill, New York, 1962).
[68] Pacejka, H. B., The Tire as a Vehicle Component (Tire In-Plane Dynamics), in Me
chanics of Pneumatic Tires, NBS Monograph 122, S. K. Clark, Ed., p. 741 (U.S.
Government Printing Office, Washington, D. C., 1971).
[69] Potts, G. R., Application of Holography to the Study of Tire Vibration, Tire Science
and Technology 1, 255-266 (1973).
I] Potts, G. R., Firestone Tire & Rubber Co. Research Report (1972) (unpublished).
I] Purdy, J. F., Mathematics Underlying the Design of Pneumatic Tires (Edwards Broth
ers, Ann Arbor, Michigan, 1963).
[72] Gough, V. E., Barson, C. W., Gough, S. W. and Bennett, W. D., Tyre Uniformity Grad
ing Machine, The Engineer 213, 731-741 (1962); SAE Paper No. 322A, (1961).
3] Pottinger, M. G., Ply Steer in Radial Carcass Tires, SAE Paper No. 760731, (1976).
'4] Gough, V. E. and Barson, C. W., Improvements in or Relating to Pneumatic Tyres,
British Patent 972,717 (1963).
[75] Jones, F. B., Breaker Strip for Pneumatic Tires, United States Patent 3,5 16,468 (1970).
Pottinger, M. G., Tire With No Ply Steer Belt, United States Patent 3,945,422 (1976).
Longmore, D. K. and Goodall, J. R., Elastic Behavior of Radial Tire Breakers, Pro
ceedings International Rubber Conference (Brighton), pp. 389-404 (Universities
Press, Belfast, 1967).
CORD REINFORCED RUBBER 179

3.2. Loss Characteristics of Cord Rubber Composites


Introduction
The loss which occurs in a cord-rubber composite may be visualized
most simply by considering a typical stress-strain curve of such a material,
as shown in Fig. 3.26. It is seen that as the stress is cycled from point A to
point B there is a corresponding change in strain, but that on the unload
ing side of the curve the stress-strain relation is not the same as on the
loading side. The shaded area between the curves AB represents the en
ergy lost as the specimen is cycled from point A to point B and return.
This means that when a cord-rubber composite material is cyclicly
stressed, work is done on each stress cycle, and this work appears in the
form of heat, which in turn must raise the temperature of the composite or
be dissipated to the surroundings.
Mathematical Basis
The mathematical basis for the behavior of "lossy" materials is due to
Boltzmann [1] who postulated that for all real solids the stress strain rela
tions were functions not only of instantaneous values but also of the com
plete stress and deformation history of the material. This introduces time
effects into the relation. For linear viscoelastic materials one may use the
principle of superposition; namely, the stress at any time resulting from a
sequence of separate strain histories applied at earlier times is a sum of the
stresses that would have been produced had the separate strain appli
cations occurred individually.
Using the same notation as Kolsky [2], this may be illustrated by assum
ing that a uniaxial strain є1 was applied at the time /, and then held con
stant, i.e. a step strain history. At some later time t this strain has caused a
stress σ1(t — t 1). Similarly a second step strain history є2 beginning at time
t2 and held constant will at time t produce a stress σ2(t — {-,). This is illus
trated in Fig. 3.27. Super-position postulates that if the strain history con
sisted of these two strains acting together then the total stress induced in
the specimen at time / would be
oCO-o.C-O + ^O-O. (3-37)
The time dependent relationship between stress and a step strain history
is called the stress relaxation function and is often denoted by the symbol
E(t). For example, if a step strain e is applied at time t = 0, then stress at
time t is given by єE(t). The stress relaxation function has the general time
decaying form shown in Fig. 3.27a. It decays to some non-zero value. The
stress relaxation function is a material property for visco-elastic materials
just as Young's modulus is for elastic materials.
When the strain history consists of a series of N step strain increments
€„, individually applied at time / = /„ the stress σ(t) observed at a time t is
given by

o(0 =£%„£(/ -*„). (3.38)


180 MECHANICS OF PNEUMATIC TIRES

STRAIN

FIGURE 3.26 Typical stress strain curve of a material

An arbitrary strain history can be regarded as the superposition of in


finitely many step strain history increments. In this case Eq. 3.38 extends
to Eq. 3.39.

E(t - T,)</C, (3.39)


i
or

(3.39b)

In the form of Eq. 3.39b this is known as Boltzmann's superposition prin


ciple. A similar treatment could be constructed for an element which un
dergoes a given state of stress as opposed to a state of strain. If a stress σ =
1 is applied as a constant value at time / = (.), i.e. a step stress history, the
strain at any later time t is given by a creep function denoted by /(/). A
similar line of reasoning to that followed above leads to the expression for
strain as a function of time in the form of Eq. 3.40.

(3.40)

DUE TO DUE TO

CO
LJ
o:

(a) (b)
FIGURE 3.27 Principle of superposition of time decaying stresses
CORD REINFORCED RUBBER 181

Note that the stress or strain components have not been specified here but
may be either extensional or shear. The creep function J(t) is also a mate
rial property. If E(t) is known, J(t) can be computed and vice versa.

Complex Modulus
The linearity associated with the assumptions leading to Eqs. (3.3b) or
(3.4) allows one to conclude that the imposition of a sinusoidal stress on a
linear viscoelastic specimen will result in a sinusoidal strain, but one
which may not be in phase with the imposed stress. A similar inverse con
clusion can be reached concerning the imposition of sinusoidal strains,
where the resulting stress states may not be in phase with imposed defor
mations. The consequences of this are that losses will be generated in ma
terials when the resulting phase lag is not zero. If such a solid is subjected
to a sinusoidal oscillating stress, the corresponding strain is also sinusoi-
dally oscillating but in general will lag behind the stress cycle. This may
be conveniently expressed by allowing the imposed stress to lead the re
sulting strain by a phase angle δ as shown in Eqs. (3.41) and (3.42).
a •• OoSin(urf + 8) = (a0cos6)sin ut + (<JosinS)cos ut (3.41)
(3.42)
where
a = strain
e = strain
S - phase angle.
Equations 3.41 and 3.42 may be most easily visualized by the diagram of
Fig. 3.28, where a pair of rotating vectors may be used to represent the
stress and strain components under sinusoidal oscillation.
Thus we may write
a = £"eo sin ut + E"^ cos ut (3.43)
where

E'- cosS

E" = -s- sinS


«0

FIGURE 3.28 Schematic representation of stress and strain under harmonic variation
182 MECHANICS OF PNEUMATIC TIRES

which represent the real and imaginary parts of the total complex modulus
as shown in Eq. 3.44.
iE" (3.44)
E* is also a material property and can be related to E(t) or J(t).
The representation of Eq. 3.43 may be thought of as two components
perpendicular to one another. This may be used to define a complex mod
ulus, one part always at right angles to the other, such as shown in Fig.
3.29. From Fig. 3.29 it is seen that the definition of the phase angle com
monly found in vibration theory may be represented by Eq. (3.9), thus re
lating real and elastic modulii to typical vibration response. The modulus
£" is called the storage modulus since it defines the elastic part of the mod
ulus. The imaginary part E" controls the loss characteristics and is called
the loss modulus. It is related to the phase angle by the relation
E"
(3.45)

This may be seen by calculating the energy loss per cycle as shown in Eq.
3.46.
Energy Loss/Cycle
//• 2-/» d€ f *•*•
ode = a — dt = a e2, / (E' sin ut cos at
_vcle 'O Ot Jo

+ E" cos2 ut)dt (3.46)


Eq. 3.46 is the origin of two expressions commonly used in explaining
the nature of losses in rubber and cord-rubber composite products. In the
first case, the energy loss per cycle at constant strain amplitude may be
taken directly from Eq. 3.46 as
Energy Loss/Cycle = irE" e2,. (3.47)
On the other hand, there exist many situations in which fixed loads are im
posed on a product. In this case, it is desirable to rewrite Eq. 3.47 in terms
of stress amplitude. The amplitudes are related according to

•-r

FIGURE 3.29 Real and imaginary components of modulus


CORD REINFORCED RUBBER 183

where |£*| = V(£')2 + (E'J


so that at constant stress level σ0

Energy Loss/Cycle = it -^—5 o02. (3.48)

Eqs. 3.47 and 3.48 have been used to explain the different roles of tread
and carcass in the power loss of a pneumatic tire. Since this only depends
on E", then the quantity E" may be viewed as a measure of the loss char
acteristics of a particular material.
Physical Models
It is common in the literature to present data on the loss characteristics
of materials in terms of the loss modulus E", or in terms of the corre
sponding tan δ, but usually as functions of temperature or frequency or
both, since most materials used in the construction of pneumatic tires are
quite sensitive to those parameters. This dependence on temperature and
frequency of real polymeric materials unfortunately does not parallel the
frequency dependence observed on the simpler mathematical models
which often are used to demonstrate the effects of viscoelastic character
istics in materials. Most of these models are not adequate representations
of real materials in any sense but are useful primarily for purposes of illus
trative convenience. One of the simplest mathematical models is the Kel
vin-Voigt model illustrated in Fig. 3.30. The constitutive equation for
such a solid is given by Eq. (13)

«-& + T,-^- (3.49)

Using Eq. 3.42 and Eq. 3.49 the real and imaginary parts of the complex
modulus are, by Eq. 3.43
£" = £ E"-Tfl3.

Another simple viscoelastic material is the Maxwell solid composed of a


spring and dash-pot arranged in series as shown in Fig. 3.31. This material
exhibits the stress strain relation shown in Eq. 3.50.

-£"- <3-50>
By using Eqs. 3.42 and 3.43 in Eq. 3.50, this gives

/ ELASTIC SPRING OF RATE E

I-VVS,-1
FORCE-*-1 -•• FORCE

—5—I
VISCOUS DAMPER OF RATE 77

FIGURE 3.30 Mechanical model of Kehin- Voigt solid


184 MECHANICS OF PNEUMATIC TIRES

FORCE- FORCE

ELASTIC SPRING VISCOUS DAMPER


OF RATE E RATE 77

FIGURE 3.31 Mechanical model of Maxwell material

E"
E2 + E2

Finally, a slightly more realistic model may be achieved by proper com


bination of two springs and one viscous damper in the form shown in Fig.
3.32. While one could attempt to describe the response of real materials by
using combinations of these various models, including the standard linear
solid, this process can become cumbersome mathematically and it is more
common in the literature to simply present the results of experimental ob
servation in the form of tan δ or E" as functions of frequency, temper
ature, amplitude or other material characteristics, all of which are signifi
cant variables in cord rubber composites.
Measurement Methods
The measurement of loss characteristics of polymeric materials is com
plicated by the extremely wide scale of frequency which must be covered.
This has been well illustrated by Ward [3] and Fig. 3.33 is taken from his
recent book. This shows that different experimental methods must be used
over the extremely broad band of frequency necessary to completely de
scribe the mechanical behavior of a polymer. Some of the more commonly
used methods encompassing the frequency range met in pneumatic tires
are listed below. These are taken from the excellent review of experimen
tal methods for determining losses in composite materials recently pub
lished by Bert and Clary [4].
(a) Free Vibration
The decay characteristics of elements in free vibration, usually beams,
may be used as a measure of the damping characteristics of the materi
als.
(b) Rotating Beam Deflection
A rotating shaft may be observed to deflect at an angle with respect to
its gravitational force or applied load. Measurement of this angle may
be used as a measure of the damping characteristics of the material.

FORCE- FORCE

FIGURE 3.32 Mechanical model of standard linear solid


CORD REINFORCED RUBBER 185

RESONANCE
METHODS
! FORCED f—1 WAVE
I VIBRATIONS i I PROPAGATION
STRESS RELAX. |
CREEP !
PENDULUM
VIBRATIONS
1 i I 1 i I i I 1 1 1
IO"8 10~6 10~4 10~* \ !02 I04 tO6 108
FREQUENCY (Hz)
FIGURE 3.33 Approximate frequency scales for different experimental techniques

(c) Sinusoidal Forced Excitation


This technique can be carried out at any frequency but generally is re
stricted to frequencies away from the resonant characteristics of the sys
tem. This type of measurement would be exemplified by that of Ref. 5,
where the direct stress-strain characteristics of the material are plotted
and used as a measure of hysteresis.
(d) Forced Resonant Vibration
In this method, a specimen is excited at frequencies in the neighborhood
of the resonant point. The damping and stiffness characteristics may be
obtained by observing the half power point frequencies, although any
other two observations of frequency and response amplitude would suf
fice.
(e) Continuous Wave Propagation Techniques
These have not been in extensive use for composite materials but are of
ten used in characterizing polymeric filaments, as in the widely used
Vibron instrument.
(f) Pulse Propagation Techniques
Such techniques have been widely used in characterizing the elastic
constants of rigid composites, but so far have not found wide appli
cation in the rubber industry.
Some of these techniques have been used by Kolsky [2] to obtain the
loss characteristics of various common polymeric materials, and to com
pare these with predictions based on the simple models exemplified by the
Kelvin-Voight material, the Maxwell material and the standard linear
solid. These comparisons are illustrated in Fig. 3.34. They show that the
Kelvin-Voight material exhibits increasing tan δ with frequency, the Max
well model the inverse, and the standard linear solid a positive slope at
low frequency, a maximum point, and a negative slope at high frequency.
None of the measured values of common polymers agree with the
simple models. In order to obtain agreement between models and real
data a spectrum of simple models must be postulated. In fact, as Kolsky
[2] points out, the JjesLfit is obtained by assuming that tan δ is a constant
value independent of frequency.
186 MECHANICS OF PNEUMATIC TIRES

A POLYETHELENE
B EBONITE
C PMM

/KELVIN
MATERIAL

CO

p9 0
§-.

-2
-3 MAXWELL
SOLID
-4
-5
-5-4-3-2-1 01 2345 €
LOG,0 FREQUENCY

FIGURE 3.34 Tan S vs. frequency for a variety of materials

Loss Properties of Cord and Rubber


The hysteretic or dissipative characteristics of the constituent materials"1"?
in a cord rubber assembly have been treated extensively in the literature. /
For example, a number of experimental studies have been published on
the loss characteristics of tire cord materials. To a great extent these have
used the concept that the tire cord is a linear viscoelastic material and as
such could be adequately described by measurement of its storage and loss
modulus values. Early measurement of this type were reported by Lyons
[6] and by Lyons and Prettyman [7]. Later review work by Kemnitz [5]
and by Lehmicke [8] also uses the viscoelastic concept to describe loss of a
tire cord. Much of this work has been carried out using the Vibron ma
chine, which utilizes the steady state dynamic characteristics of a cord or
filament in tension to measure the real and imaginary part of the total
modulus of the material. Table 3.1 1 is extracted from the work of Lyons
[6]. This data is similar but somewhat smaller than comparable tire cord
data taken on dipped cords in the recent tire hysteresis study of WUlett [9].
Willett's data are given in Table 3.12.
It is generally recognized that a substantial fraction of the loss of a typi-

TABLE 3.11 Range oftan 8 overfrequency spreadfrom Lyons [6]


Sample Description Frequency Range, Hz Range of tan £
Steel wire 6 mil 128-195 0.0027-0.0045 ^]
Fiberglas cord 232-359 0.0262-0.0298
1 100/2 Rayon 99-160 0.0297-0.0310
Nylon cord 79-127 0.0262-0.0287J

'-. 5
5/o
CORD REINFORCED RUBBER 187

TABLE 3.12 Loss characteristics oftire cord material* (dipped cord))after Willett [9].
Sample Description E",108Nm-2»
1 100/2 Rayon 1.70 0.16 0.052
840/2 Nylon 0.82 0.30 . 0.050
' See Eqs (3. 10) and (3.12)

he ascribed to its Coulomb interfilament frictional char-


actenstics, which may not be well represented by a loss modulus of the
type commonly associated with a viscoelastic material. Nevertheless the
presentation of the data in terms of tan δ is a convenient device and is al
most universally used.
/-» Insofar as is known, no similar work on the bending hysteresis of tire
L cords has been published.
There has been substantial amount of material published in the litera
ture on the loss characteristics of various rubber compounds under differ
ent temperatures and test conditions. Typical of such recent work is the re
view paper of Ulmer, Hess, and Chirico [10]. Here tan δ is presented asji^
function of temperature and carbon black loading in an oil extended SBR
stock. Their data are given in Figr3.35. From this data it isr seen that tan?
-for the rubber component alone is sensitive not only to the carbon black
loading but even more sensitive to the temperature at which the property
is measured.
Willett [9] has used the loss characteristics of actual tires in order to
compute the influence of different tire components on tire hysteresis and
heating. The results of his measurements on these passenger car tires are
given in Tables 3.13 and 3.14.

10% STRAIN
0.14 PHR <t>
o GUM
i>
6
0.12 o 'IS 0.0519
o 20 0.0693
A 35 0.1122

j 0.10

0.08

0.06

0.04
V
^ 0.02

20 40 60 80 100 120
TEMPERATURE (°C)

FIGURE 3.35 Tan S as a function of temperature and volume fraction ofN-220 and XI-990
in an OESBR stock
188 MECHANICS OF PNEUMATIC TIRES
0.35

SBR FftSSENGER TREADS 8


0.30 COMMERCIAL RETREAD
STOCKS

0.25 SBR TRUCK TREADS


BELT COVER STOCKS
0.20
HIGH QUALITY AUTOMOBLE
MATS
0.15 OFF-THE-ROAD TREADS (SBR)
SHOE SOLES
GASKETS AND SEALS
0.10
PASSENGER TIRE >

¥ 005
CARCASS STOCKS

RUBBERjIANDS
I ' I I I
10 20 25 30
E1 x 10" ?, dynes/cm at 70°C
FIGURE 3.36 Stiffness-hysteresis chart for nine types of commercial rubber compounds

An additional method of presenting such data was proposed by Stu-


de baker and Beatty [11] in the form of a plot of tanδ vs the real part of the
complex modulus. On such a plot it is possible to identify the areas in
which ^commercial products lie, and one of the composite diagrams from
their recent paper is given as Fig. 3.36. From this diagram it is possible to
see that in general, tread stocks tend to lie in the region of relatively high
tan δ with moderate elastic moduluSj, To some extent this comes about due
to the desire to minimize abrasion and wear by means of carbon black ad
ditions to the tread stock, which tend to raise the loss component. Carcass
stocks, on the other hand, are compounded to operate in region of niuch
lower tan δ as shown in Fig. 3.36, in part because they are confined to in
ternal regions of the tire where heat transfer characteristics are much
worse so that the temperature rise associated with losses would tend to be
much greater there than in the tread.
AUjof thfise studies show that the loss modulus of typical rubber com
posites can be strongly affected by carbon black loading, by strain ampli-
lude'|"by oil extension and by temperature. This latter effect is of consid
erable importance in actual tire service where the hysteretic loss results in
increased tire temperatures. These in turn result in decreased hysteretic
loss and the system is self stabilizing. Were this not the case, and were the

TABLE 3.13 Properties ofply rubber in test tires after Willett f)].
Tire No. E"(10*Nm-2) lamS
0.820 0.363 0.173 ~
0.626 0.491 0.175
0.929 0.269 0.158 J
CORD REINFORCED RUBBER 189

TABLE 3.14 Properties of tread rubber in test tires after Willett [9].
Tire No. E" (10*Nm-2) 12"/(E'hlO-7m*N-') tanS
1 0.235 0.710 0.129 -
2 0.266 0.880 0.153
3 0.306 0.928 0.169 \
4 0.352 0.900 0.178
5 0.400 0.650 0.161 J

hysteretic loss to rise with increasing temperature, then the system might
very well become unstable.
Loss Properties of Composites
There is at this time very little published information on the general
characteristics of the assembled cord rubber composite itself. Partly this is
due to the fact that design judgements in the past have been made primar
ily on the basis of the loss characteristics of the components used in the
tire, namely the cord and rubber. There is now suspicion that structure
also plays an important role here, so That it is not sufficient simply to rely
on characteristics of the constituent materials in order to be able to esti
mate the loss cliaracteristics bFthe composTfeT
Hash in [12, 13] has developed a theoretical framework in which the
time dependent properties of viscoelastic composites may be deduced
from knowledge of their elastic characteristics by use of the correspon
dence principle used so widely in viscoelasticity. This process devised by
Hashin shows that "the effective complex moduli! of a viscoelastic hetero
geneous specimen can be found by replacement oTthe phase elastic modu-
ui by phase complex modulii in the expressions for the effective elastic
modulii of an associated heterogeneous elastic specimen with identical
phase geometry."
In theory this should allow the use of the Halpin-Tsai equations (see
Section 3.1) to be used in conjucntion with the known viscoelastic charac
teristics of the constituent materials in order to predict the complex modu
lus of a single sheet of calendered fabric, and from that basis to use equa
tions such as Eqs. (3.2-3.5) of Sec. 3.1 to predict the complex modulus of
the assembled composites.
Unfortunately the Hashin principle requires that the constituent mate
rial phases be linearly viscoelastic and homogeneous. While this condition
is approximated by the rubber component of a typical cord-rubber system,
the cord component may not satisfy this requirement. Very little informa
tion is yet available on the adequacy of the Hashin principle in predicting
composite complex modulii from known constituent properties for the
commercially important cord-rubber combinations, even on a relatively
low frequency basis.
It is, however, in the area of higher frequency effects that the Hashin
principle may suffer its greatest departure from reality, since it is formally
true only for representative volume elements small enough so that stress
and strain can be treated as statistical averages over these areas. For
higher frequency motions where wave length is comparable to cord spac
ing or diameter the theory is not valid.
Insofar as is known the experimental evidence published on the ade
quacy of the Hashin principle in predicting the complex modulii of com
190 MECHANICS OF PNEUMATIC TIRES

poshes is all taken with monofilament fibrous reinforcement networks,


and not with twisted filament cord assemblies such as in tire cord. [18].
Of additional importance in explaining the observed loss characteristics
of composites are thermorheological effects. It is observed that cord-rub
ber composites often become hot in service. This heating is a result of in
ternal losses in the specimen or article, and the final temperature is depen
dent not only on material properties but also on specimen geometry,
dimension, stress amplitude and frequency. Due to the complexity of this
problem few attempts have been made in the literature to deal with it. The
recent paper of Holownia [15] treats the specific case of a cylindrical block
of rubber under cyclic compression. He shows that temperature build-up
can be predicted from conventional heat conduction and elasticity con
cepts, but that numerical solutions must be employed for even this simple
geometry. His results show that temperature build-up depends on speci
men geometry, a conclusion which merits further study. Other recent pa
pers dealing with cyclic heating are those of Huang and Lee [16] on forced
motion of rods made of a rocket propellent, and Tauchert [17] on the tor
sion of a polyethylene rod.
There have been few reports of experimental work in the literature on
the assembled cord rubber composite. Skelton [14] has recently suggested
a method of using ribbon-like specimens to obtain the bending hysteresis
characteristics of cord rubber composites. This is illustrated in Figs. 3.37
and 3.38 which show the experimental apparatus and a typical load de
flection curve taken from it. Data obtained from such tests is summarized
in Fig. 3.39, where the loss characteristics in bending are illustrated for
rubber specimens reinforced with various types of tire cord materials.
There has been very little other direct work published on cord rubber
composites so far, although it is anticipated that substantial activity in this
area will be forthcoming soon due to the considerable interest in the re
duction of rolling loss in pneumatic tires. However, quite aside from this
point, the measurement of composite hysteretic characteristics is impor
tant since these characteristics determine the operating temperatures of
products which are subjected to cyclic stress. An excellent review of the

TO
INSTRON
LOAD CELL

u
TO INSTRON
CROSSHEAD

FIGURE 3.37 Bending hysteresis apparatus


CORD REINFORCED RUBBER 191

FIGURE 3.38 Bending moment-curvature relation

STEEL

RAYON
GLASS
NYLON POLYESTER
RUBBER ALONE

0.5 1.0 1.5 2.0 2.5


MAXIMUM CURVATURE (cm'1 )

FIGURE 3.39 Variation of bending hysteresis loss/cycle with maximum curvature for various
specimens

current status of this area has been given by Schapery [18]. In view of the
large degradation of both polymeric cord and rubber strength and fatigue
life with increased temperature, the role of hysteresis is extremely impor
tant in design since it controls the temperature by determining the amount
of energy which must be dissipated from the product. This alone is suf
ficient reason for greater efforts to understand the role of composite struc
tures in the hysteretic loss process.

References
[1] Bolumann, L , Pogg. Ann. Phys. Chem., v. 7, 624 (1876).
[2] Kolsky, H., "The Role of Experiment in the Development of Solid Mechanics—Some
Examples," Advances in Applied Mechanics, v. 4, Ed. by C. S. Yih., Pergamon Press,
New York, 1976.
192 MECHANICS OF PNEUMATIC TIRES

[3] Ward, I. M., "Mechanical Properties of Solid Polymers," John Wiley & Sons, New York,
1971.
[4] Bert, C. W. and R. R. Clary, "Evaluation of Experimental Methods for Determining
Dynamic Stiffness and Damping of Composite Materials," A.S.T.M. Special Techni
cal Publication 546, American Soc. for Testing and Materials, Phil., Pa., 1974.
Kemnitz, G., "Untersuchung der Stauch-Ermuclung and der Dampfungs-Erscheinu-
ngen on Reifen-cord," Kautshuk u. Gummi, v. 12, p. WT270, 1959.
) Lyons, W. J., "Dynamic Properties of Filaments, Yarns, and Cords at Sonic Frequen-
' ties," Textile Res. Journ., v. 19, 123, 1949. •
[7] Lyons, W. J. and I. Prettyman, J. Appl. Physics, v. 19, 473, 1948.
[8] Lehmicke, D. J., "Effect of Twist and Treating Conditions on the Physical Properties of
PET Tirecord," Div. Rubber Chem., A. C. S., Chicago, Oct. 1970.
[9] Willett, P. R., "Hysteretic Losses in Rolling Tires," Rubber Chem. Tech., v. 46, n. 2,
June 1973, p. 425.
nm Ulmer, J. D., W. M. Hess, and V. E. Chirico, "The Effect of Carbon Black on Rubber
%/ Hysteresis," Rubber Chem. Tech., v. 47, n. 4, Sept. 1974.
(fill Studebaker, M. L. and J. R. Beatty, "Effects of Compounding on Dynamic Mechanical
VJ Properties of Rubber," Rubber Chem. Tech., v. 47, n. 4, Sept. 1974, p. 803.
[12] Hashin, Z., "Complex Moduli of Viscoelastic Composites-I. General Theory and Appli
cation to Particulate Composites," Int. Jour. Solids Structures, 1970, v. 6, p. 539.
[13] Hashin, Z., "Viscoelastic Fiber Reinforced Materials," AIAA Jour., v. 4, n. 8, August
1966, p. 1411.
Skelton, J. R., "Bending Hysteresis Losses in Cord-Rubber Composites," Rubber
Chem. Tech., v. 47, n. 2, June 1974.
ffi5] Holownia, B. P., 'Temperature Buildup in Bonded Rubber Blocks Due to Hysteresis,"
^~S Rubber Chem. Tech., v. 50, n. 1, March-April 1977, p. 187.
[16] Huang, N. C. and E. H. Lee, "Thermomechanical Coupling Behavior of Viscoelastic
Rods Subjected to Cyclic Loading," Jour. Appl. Mech., v. 34, n. 1, March 1967, p.
127.
[17] Tauchert, T. R., "The Temperature Generated During Torsional Oscillations of Poly
ethylene Rods," Int. J. Engng. Sci., v. 5, 1967, p. 353.
[18] Schapery, R. A., "Viscoelastic Behavior and Analysis of Composite Materials," Me
chanics and Materials Research Center, Texas A. & M. University, College Station,
Texas, Aug. 1972.
CORD REINFORCED RUBBER 193

33. Failure Mechanisms of Cord Reinforced Rubber


The failure mechanisms of cord reinforced rubber are extremely com
plex and have recently been the subject of considerable active research.
For purposes of application to pneumatic tires, most of the emphasis lies
on the fatigue failure characteristics of these materials, since service re
quirements demand that a pneumatic tire undergo a very large number of
loading cycles. This inevitably forces the tire designer to proportion the
carcass textile reinforcement in such a way that ultimate static strength is
many times the inflation pressure or the applied load. Large static factors
of safety are the rule in tire design.
When ultimate tensile strength is to be calculated, most approximations
consider that cord failure results when cord loads reach a tensile value
equal to that which they exhibit in simple direct tension tests. In other
words, ultimate strength of cord reinforced rubbers is equivalent to cord
strength uninfluenced by the surrounding rubber, so that the rubber nei
ther increases nor decreases the ultimate tensile strength of the cord. To
that extent the cord-rubber system may be thought of as uncoupled in re
gard to ultimate strength. For example, bursting of a tire due to inflation
would be predicted on a cord tension basis without regard to presence of
the rubber.
The more common loading environment involves the fatigue character
istics of cord reinforced rubber, and here the situation is much more com
plex. First of all, it is necessary to recognize that the complete cord rein
forced rubber system consists of three phases: first, the textile cord itself,
secondly, the adhesive which is applied to the cord surface in order to pro
mote adhesion to the rubber, and finally the surrounding rubber matrix.
Details of the rubber and cord characteristics are given in previous chap
ters of this volume, but it should be noted that the cord is made of a multi
plicity of filaments which are first twisted to form a strand, followed by the
twisting in opposite directions of two or more strands to form a cord. For
purposes of pneumatic tire construction two and three strand cords are the
most commonly used.
Since many service failures seem to occur as a result of fatigue, a great
deal of attention has been directed to mechanisms of tire fatigue failure. In
order to describe such mechanisms, it is probably best to do so in some
systematic order roughly parallelling that which would occur in an actual
tire. For such a description, one may draw upon the work of Eccher [1],
Patterson and Anderson [2], Patterson [3], Uzina and Basin [4], Butter-
worth, Davis, and Platt [5]. A number of other authors and references will
be cited in addition, but the basic mechanism of fatigue failure as it is now
understood is well delineated by these authors.
First of all, two primary concepts regarding fatigue failure should be
stated clearly:
a. For cord reinforced rubber, the so-called fatigue failure is, in fact, a
system failure in which the failure is initiated at or very close to the cord-
adhesive-rubber interface.
b. A normal fatigue process, such as occurs in metals, does not exist for
194 MECHANICS OF PNEUMATIC TIRES

TABLE 3.15. Depth to which latex-resorcinal-formaldehyde compositions penetrate various


cords when dipped

Viscose Nylon Cotton


Cord diam. (mils*) 35 20 35
Radial dip penetration in fiber diameters 4 3 4
Radial dip penetration (mils) 1.9 3.54 2.8
Fiber diam. (mils) 0.485 1.18 0.7
• 1 mil = 0.001 inch.

the cord-adhesive-rubber combination as a whole. However, there exist fa


tigue mechanisms both for the rubber and for the rubber to adhesive
bond, which then directly or indirectly act on the strength of the fabric.
The actual initiation of the processes which precede the fatigue failure
of a cord reinforced elastomer has been studied by Eccher [1], by use of
the Mallory tube test and by microscopic examination, and by Uzina and
Basin [4], using fluorescence analysis as well as photomicrography. Uzina
and Basin studied dipped viscose, Nylon 6, cotton and polyamide cords
under fatigue conditions and observed the following forms of initial fail
ure on tires which had been run to the point of obvious ply separation:
a. Cord separation accompanied by a considerable part of the adhesive
adhering to the rubber, coupled with some adhesive remaining on the
cord. Such a mechanism clearly indicates separation occurring within the
adhesive layer, but is cohesive in nature.
b. Separation in which a thin film of the adhesive remains attached to
the rubber, while in some areas a portion of the rubber remains with the
adhesive. This type of separation is much closer to a separation along the
adhesive-rubber boundary than type a., and can be termed a composite
separation. It is also cohesive in nature.
c. Separation occurring without traces of adhesive film being left on the
rubber or of rubber on the cord. Consequently, this separation occurs
cleanly at the adhesive-rubber interface and is purely adhesive in nature.
Some of the details of such a cord-rubber-adhesive system are presented
here. Table 3.15 gives typical penetration depths for RFL3 dipped cords.
Eccher [1] found by photomicrography that both Mallory test tubes as
well as examination of tires implied very similar types of failure initiation.
Butterworth, Davis and Platt [5] have accumulated extensive evidence
to show that the types of failures just described are promoted and acceler
ated when the cord is forced to carry compression in the pneumatic tire
structure. Their studies agree with those of Patterson [3], and both authors
report that compressive loading of typical tire cord structures in trans
parent rubber allows one to clearly observe the strain concentration effects
of compressive forces on the cord-rubber interface. In particular, the reen
trant angle caused by the twisting of the several strands together to form
the cord seems to be a point of particularly high strain concentration, and
a point which is prone to initiate adhesive system failure under compres
sive loads. Cord degradation resistance is determined by the ability of the

RFL is the common abbreviation for resorcmal-fbrmaldehyde-lalex, which is the basic tire cord adhesive. Most com
mercial adhesives use this u a basis for more sophisticated, proprietary adhesives.
CORD REINFORCED RUBBER 195

surrounding rubber matrix to control the compressive axial deformation


of the cord. Where the adhesive interface fails, this control is lost and fila
ments traversing the reentrant angle in the cord geometry tend to buckle
in an uncontrolled manner. This leads to severe relative deformation be
tween the surrounding rubber matrix and the cord structure. This relative
deformation increases in severity because, with continued cycling, the
cord geometry opens up (reduction in cord packing factor) leading to in
creased filament buckling and an increased interaction with the matrix. A
reduction in cord packing factor leads to the induction of higher tensile
strains in filament lengths having a low angle of inclination to the cord
axis—mainly filaments on the external surface of the cord. Test data in
general indicates that cord fatigue is greatly dependent upon the presence
of compressive stresses in the cord, as reported by Illingworth [7].
Following the initial failure somewhere in the cord-adhesive-rubber
system, present evidence points toward the occurrence of increased rela
tive motion between the cord and rubber, as well as among the cord fila
ments themselves. This may result in chafing or rubbing of the cord fila
ments against one another, with the consequent possibility of initiating
sites of microscopic mechanical damage on the surface of the individual
filaments. Patterson [3] has studied this in some detail and concludes that
the mechanism of filament rupture in the fatigue of individual nylon
monofilaments is directly traceable to the presence of some small surface
defect. Such a defect serves as an initiator of fatigue failure, which then
propagates due to alternating bending of the monofilament. Patterson
found that if the imperfection occurs at the point of maximum compres
sion in bending, the fatigue life of a monofilament was greatly decreased.
Simple bending seems to account for the mechanism of failure since the
appearance of broken filaments taken from pneumatic tires and broken
filaments fatigued in bending are essentially identical. Figure 3.40 shows
an enlargement of a filament taken from a tire after some running.
This same problem has recently been studied by Platt and Butterworth.
They conclude that the creation of a filament buckle and the location of
the buckle angles is determined by the position of the element of filament,
its degree of association with neighboring filaments and triaxial stress state
to which it is subjected. They conclude that it is unlikely that all filament

FIGURE 3.40 Bias ruptured filament.


1% MECHANICS OF PNEUMATIC TIRES

FIGURE 3.41a Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Failure initiated on inside of bend flex zone. Nylon monofilament whose diameter is equivalent to a standard 2 ply nylon
cord - mils.

failures are initiated by a defect of a filament leading to a buckle failure.


They observed that when monofi laments of diameter equivalent to a con
ventional tire cord are subjected to simulated tire use, the monofi lament
tends to buckle in fashion similar in shape and character to buckles noted
in individual filaments taken from fatigued tire cords. Such buckles are il
lustrated in figure 3.41. In particular, the two angle buckle shown there
failed initially on the compression side of the buckle but ultimately failed
in tension.
The process of tire degradation therefore apparently involves failure

FIGURE 3.41b Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Two bend buckle of 20 nil nylon mononlament.
CORD REINFORCED RUBBER 197

FIGURE 3.41c Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Partially filled nylon lire cord. Degraded in rubber on Goodrich ten machine
3 percent tension cyclic axial
IS percent compression cord strain
Note inception of filament buckle failure at cord ply line

FIGURE 3.41d Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Failure sequence in two-ply dipped and stretched nylon tire cord.
198 MECHANICS OF PNEUMATIC TIRES

FIGURE 3.4 le Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
20 mil nylon monofit. failure in rubber, induced by tensile and compreuive strain cycling.

first at the surface of the cord, by one of the three forms proposed by
Uzina and Basin, allowing detachment of the cord from the surrounding
rubber. This is followed, according to Butterworth, Davis, and Platt [5], by
mechanical working of the cord leading to a reversion of its tensile proper
ties to those of the undipped, unstretched cord. Failure of individual fila
ments next occurs, in a sequential fashion, finally leading to total cord fail
ure. This hypothesis seems to be in accord with the experimental evidence
obtained by studying the characteristics of tires which have been run on
both road and test wheels, since most observations have found that tire
cords which have lost a fraction of their strength due to monofilament
fracture in fatigue also exhibit reduced adhesive bond strength to the sur
rounding rubber, as measured by strip-out tests.
The process of cord degradation goes forward by means of individual
monofilament fatigue fracture. The nature of this process is clearly statisti
cal and is not well understood. Patterson and Anderson [2] carried out an
extensive study of the residual strength of nylon tire cords in tires which
had been run on test wheels, while similar studies on rayon cords in high
way service have been carried out by Klein, Platt, and Hamburger [8].
Both authors conclude that cord strength does not linearly degrade with
mileage. As a matter of fact, it is this particular area which represents one
of the substantially unknown regions in the process of tire cord degrada
tion, since both these studies indicate that after an initial, relatively rapid
loss of cord strength there follows a long period in which the rate of loss is
rather small. Typical data are shown in figures 3.42 and 3.43. The reason
CORD REINFORCED RUBBER 199

FIGURE 3.41f Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Partimlly filled nylon tire cord.

for this initial loss is as yet uncertain, but it has been postulated that it is
due to physical and geometric rearrangement. Even in fatigue-failed tires,
cords adjacent to the failure are in general of normal strength. It must be
concluded that the cord failure process, once initiated, is somehow cata
strophic and runs a rapid course, similar to the effects found in propaga
tion of a fatigue crack in metals.
There are substantial differences between cord failure rates in different
parts of the tire. Patterson and Anderson report the most critical regions to
be at the shoulder and shortly below the turn-up region, as shown in fig
ure 3.44, and further depends on the direction of the cord with respect to
the tire direction of motion, as previously shown in figure 3.42.
Eccher [1] has discussed in some detail the use of various tests, and par
ticularly the Mallory tube test, as a mechanism for simulating the behavior
of the tire materials in service insofar as such fatigue characteristics are
concerned. He concludes that the Mallory test, which uses a tube bent
through a 90° arc, held under internal pressure and rotated at high speed,
is valuable in reproducing cord-adhesive-rubber system failure similar to
200 MECHANICS OF PNEUMATIC TIRES

FIGURE 3.41g Filament failure modes. (Courtesy of Fabric Research Laboratories, Inc.)
Cord crou section photomicrographs Failure sequence A-F.

14
13

12

AVERAGE "
CORD ..
BREAKING lo TRAILING
STRENGTH N SHOULDER
KGMS
*%i_ /

LEADING SHOULDER
i i i i i i
"0 2 4 6 8 10 12 M 16 18
MILES ON TEST WHEELJHOUSANDS

FIGURE 3.42 Average breaking strength of sections offirst-ply cords.


CORD REINFORCED RUBBER 201

100

Viscose 8. 1100/2 '


Viscose A, 1100/2
Viscose B , 1650/2
Viscose A, 1650/2
Nylon . 640/2

20 40 60 80 100
THOUSANDS OF MILES

FIGURE 3.43 Percent residual cord strength vs. tire mileage.

those observed in tires. A sketch of the Mallory apparatus is shown in fig


ure 3.45. A number of other test devices have been designed to induce
cord fatigue, and since this is an area related but not particularly pertinent
to tire mechanics, the reader is again referred to Eccher [1] for a more
thorough discussion of various test methods.
Once a cord has failed, the loss of strength throws additional load to
neighboring cords in such a way that under favorable conditions an adhe
sive separation of one of the types previously mentioned may begin in the
neighboring cord, followed by individual filament rupture and eventual
cord tensile fracture. Under sufficiently severe loads this process can result
in the propagation of a failure.
Finally, it might be of some brief interest to indicate the magnitude of
the problem of fatigue failure of tire materials. This can only be estimated
roughly, but the data of Starks [6], taken on the British Motorway M.1 in

REGION OF HIGH STRENGTH


LOSS AT SHOULDER

REGION OF HIGH STRENGTH


LOSS AT END OF FLIPPER

FIGURE 3.44 Locations of maximum cord damage in tire after flexing test.
202 MECHANICS OF PNEUMATIC TIRES

COUNTER

REGULATED AIR
PRESSURE SUPPLY

FIGURE 3.45 Sketch of Mallory lube lest.

1962-63, showed that the percentage of total tire failures was caused al
most exactly half by puncture and half by burst, the latter representing a
fatigue failure of some sort or another. Therefore, the question of the fa
tigue of cord reinforced rubber is not only complex but very practical in
deed.

References
[1] Eccher, S., Typical damage in Mallory tubes and tire carcasses. Rubber Chem. Tech. 19,
299 (1966).
[2] Patterson, R. G., and Anderson, R. K .. Fatigue failure in nylon reinforced tires. Rubber
Chem. Tech. 38 (4) (Nov. 1965).
[3] Patterson, R. G., Mechanics of bias filament rupture in fatigue of nylon, Rubber Chem.
Tech. 39 (5) (Dec. 1966).
[4] U/.ina, R. V., and Basin, V E., Study of nature of failure in the system cord-adhesive
rubber, Sov. Rubber Tech. 19, 27-33 (July 1960).
[5] Butterworth, G. A. M., Davis, S. J., and Platt, M. M., The response of nylon tire cord in
rubber systems to cyclical strain (Fabric Research Laboratories, Inc., unpublished re
port).
[6] Starks, H. J. H., Tyre failures in accidents on the motorway M. 1 in 1962-63, Inst. Auto.
Asses. 17 (1), 17-25 (1966).
Illingworth, J. W., J. Textile Inst. 44, 328 (1953).
Klein, W. G., Platt, M. M., and Hamburger, W. J., Cord fatigue in fleet tested tires (Fab
ric Research Laboratories, Dedham, Mass.).
(9] Platt, M. M., and Butterworth, G. A. M. (Private communication).
Chapter 4
STRUCTURE OF THE PNEUMATIC
TIRE
V. E. Gough

4.1. General considerations 204


4.2 Pneumatic tire structure—general features 207
4.3. Flexible filament and soft matrix constructions 216
4.4. Tire construction methods 225
4.5. Cord path—practical factors determining selection 229
4.6. Calculation of cord length 231
4.7. Analysis of manufacturing methods by R vs. cos <j> charts 235
4.8. Mechanism of load carrying—infinitely flexible membrane ... 237
4.9. Mechanism of load carrying—tire structure 243
References 248

Formerly with the Dunlop Company, Birmingham, England (deceased)


to
20)
204 MECHANICS OF PNEUMATIC TIRES

4.1. General Considerations


That the function of a tire is to transmit the forces which drive, brake,
and guide the vehicle as well as to carry the load is now accepted by all—-it
is, of course, the only component of the vehicle which makes contact with
the road surface. Also, it is now being recognized that the tire not only has
to absorb local road surface irregularities over a wide range of types of
road materials but also has to provide a vibration-free motion on smooth
roads comparable with that given by a circular solid wheel on a perfectly
straight guideway. There is a conflict in these differing requirements. Flex
ibility is required for the absorption of road rugosity. Constancy of effec
tive dimensions is needed for the exacting requirements of constant axle
height, straight line motion and uniformity of effective rolling radius on a
flat, smooth road as illustrated in figure 4.1. This conflict is further height
ened by the fact that basically a pneumatic tire structure is a surface of
revolution, usually of curved cross section in a radial plane, and con
sequently is a surface of double curvature. A surface of double curvature
is, from a geometrical point of view, a nondevelopable surface and yet it
must deform to give an area of contact on a plane road surface.
These conflicting geometric and mechanical requirements demand that
a vehicle tire is either:
a. a solid tire made to a high geometrical precision from relatively low
modulus, highly elastic material capable of substantial deformation,
or
b. a gas inflated envelope with certain exacting requirements to be met.
It will be seen later that this envelope of double curvature usually is an an-
isotropic hollow structure made of flexible filaments of high modulus ma
terial such as textiles, metal or glass, embedded in and bonded to a matrix
of low modulus material such as rubber or a rubber-like polymer. The ori
entation of the filaments in this envelope has to be such as to meet certain
contrasting structural requirements. These are:
a. no appreciable change of size upon inflation,
b. an ability to envelop obstacles without sustaining damage,
c. the ability of part of this envelope to deform from a surface of
double curvature to a plane surface, and
d. enough rigidity to develop substantial forces in any direction.
It is this type of structure in which we are interested.
Inflated shell structures having surfaces of double curvature made solely
of a single isotropic material have in the past failed to meet all of the sev
eral requirements outlined above. The capability of having a long fatigue
life when continuously flexed from a surface of double curvature to a flat
STRUCTURE OF THE PNEUMATIC TIRE 205

a. b.

FIGURE 4. 1 Conflicting requirements which a tire has to meet.


(a) ability lo absorb surface irregularities—this requires flexibility
(b) constant axle height and effective rolling radius on smooth roads—ihis requires uniformity, and is most easily obtained
with a rigid wheel.

surface and back again—and at the same time having adequate structural
rigidity to carry the vehicle load and resist the drive, brake and side force
for a practical automobile, and also at the same time providing a constant
dimensional size nearly independent of the inflation pressure, have not yet
been met successfully by an isotropic thin or thick shell of materials al
though research is still being carried out.
Uninflated thin shell structures of double curvature made from high
modulus materials such as metals or the semi-rigid or rigid plastics, in
cluding all known forms of high modulus polymers, have not been practi
cal since the necessary flat contact cannot be developed without stressing
the material beyond its yield point or at least to the point where fatigue
life is very short.
The same limitations apply to open frame structures, unsupported by
gas inflation, made from a single material or a combination of materials,
as anyone who has had experience in designing spring wheel alternatives
to the pneumatic tire will agree. Such structures can, however, be designed
to meet the requirements associated with low gravitational fields, e.g.,
moon vehicles. Consider the design of a tire-like space frame consisting of
a series of steel rings of appropriate dimensions for use on an automobile
on earth, such as shown in figure 4.2. It is found that either the rings are
thin enough to be capable of deflecting the required amount but can only
carry a trivial load, or, they are stiff enough to carry a useful load but too
stiff to change their shape to a flat one without exceeding the yield point or
fatigue stress limits of the metal or other material being used. Practically
the only type of spring wheel design to get further than the drawing board
stage is a series of helical springs supporting a chain belt or flexible ring as
shown in figure 4.3. This is a sort of solid tire in which the low modulus
material is replaced by suitably fashioned high modulus material to get a
long length of metal torsion bar in the space available. Such devices have,
however, never seriously challenged even solid rubber tires, let alone the
pneumatic tire.
Consideration of these various concepts and designs may seem to be
somewhat removed from the structure of a pneumatic tire, but they have
some relevance to the fact that a pneumatic tire casing is a practical work
ing device precisely because it is comprised of high modulus flexible fila-
206 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.2 Space frame and spring wheels built of steel hoops in any configuration can be
designedfor either load requirement or deflection requirement, but not both for a practical earth
vehicle.

ments embedded in and bonded to a low modulus matrix. Such an ani-


sotropic shell structure, although in most configurations unable to carry
much load by itself, can carry substantial loads when inflated by a gas un
der pressure, and can still meet the geometric and mechanical require
ments stated earlier. One reason for this is that the ground contact pres
sure within the contact patch is primarily determined by the pressure of
the gas inflating the tire structure, so that the load carried by the tire is not
limited by the conflict of material properties in the same way as it is in the
case of the solid rubber tire. In the case of the solid tire the entire load is
carried by only a small fraction of the total tire volume, in fact, to a first
approximation it is carried by a volume made up of the tire cross-sectional
area times the length of the contact patch. To carry a greater load at no
greater tire stress requires a wider, heavier tire. The total weight of the tire
is thus approximately the ratio of the tire periphery to the contact patch
length multiplied by the weight of the material being locally stressed. In a
pneumatic tire, on the other hand, such a calculation is not pertinent since
load carrying ability is controlled by contact patch area and inflation pres
sure. Increasing inflation pressure contributes negligibly to tire weight.
Although this indicates the impracticability of designing a structure
which does not use a gas as its main supporting medium, it is possible to
produce constructions which will give a limited life and adequate handling
STRUCTURE OF THE PNEUMATIC TIRE 207

FIGURE 4.3 One form of spring wheel surviving initial tests, but which does not compete with
pneumatic tires.

properties under emergency deflations; this gives safety and the capacity
to continue to drive the vehicle to a service station.
All these points and several others lead to the fundamental importance
of the character of the structure of the pneumatic tire.
Before describing in outline the character of the pneumatic tire struc
ture it must be stated that full or complete mathematical analyses of these
structures in the form of closed functions are extremely complex because
of the highly redundant nature of the structure.
Useful analyses, usually approximate in character, of a number of as
pects of major importance are possible. This is so in spite of the fact that
the relevant mechanics problems are usually considered difficult subjects,
such as finite deformation of a nondevelopable surface of revolution, non
linear elastic and time dependent or nonlinear hysteretic properties of the
materials, and added complexity due to anisotropy of the basic structure.
The fact that the pneumatic tire is essential to the basis of twentieth cen
tury living and that innumerable forms, types, and ranges of sizes are
manufactured on a mass production basis without any noticeable com
petition from alternative devices shows that tire designers have succeeded
in producing viable and efficient product designs using the simple but the
oretical analyses or concepts which are available to them.
The objective in this chapter is to try to examine the fundamentals of
why the pneumatic tire is made from filaments embedded in a lower mod
ulus matrix and to point out the principles of the consequences of this fact,
along with the fact that the structure requires gas inflation to be an effec
tive load carrying device. It is not intended to discuss current design rules
and practice.

4.2. Pneumatic Tire Structure—General Features


A pneumatic tire has certain essential structural elements. The most im
portant is the casing or carcass made up of many flexible filaments of high
208 MECHANICS OF PNEUMATIC TIRES

modulus cord, of natural textile, synthetic polymer, glass fibre, or fine


hard drawn steel embedded in and bonded to a matrix of low modulus po
lymeric material, usually natural or synthetic rubber.
The flexible high modulus cords are usually disposed as multi-filament
layers. One, two, or more such layers of parallel cords are used according
to the design requirements. The number of layers is decided in the first in
stance by the tire type, the tire size and the inflation pressure to be used.
The first commercially successful tires were made from woven fabric as
shown in figure 4.4a, and even sheet leather was proposed in the first re
corded design. It was only when layers of parallel cords, not woven, were
introduced that tire endurance life reached a reliable level and adequate
tire performance could be taken for granted. This is shown in figure 4.4b.
Although the principle is well established that cords in a given layer must
be in one direction only, and cords in another direction must be in another
layer or ply, the parallel cords in a ply are frequently connected by light
weight wefts, see figure 4.4c. These fine wefts are not stress carrying mem
bers and serve only to hold the cords in their appropriate relationship dur
ing manufacturing processing. The need for these wefts is principally dur
ing any dipping or adhesion treatment prior to rubbering, and during the
feed stage of the rubbering process on the calender. It is essential that the
direction of the wefts are not diverted substantially from their normal di
rection of lying at right angles to the main cords of the ply during these
process stages because, during the subsequent shearing within the ply
which occurs during the shaping stage of tire manufacture, the wefts could
damage or take a position so as to cause damage to develop during tire
life. Sometimes to avoid this problem the wefts are so designed that they
fracture during the tire shaping process during manufacture.
It would appear that a continuous ring like a hose would make a suit

FlGURE 4.4 Filament arrangements which have been used successfully in pneumatic tires, at
one time or another.
(a) Woven cord (obsolete).
(b) Weftlew ...id
(c) Cord material with light wefU.
STRUCTURE OF THE PNEUMATIC TIRE 209

able tire construction, provided that the tubular tire was so attached to the
wheel on which it is mounted that it could not roll off the wheel under the
action of sideforce. While constructions of this type have been made for
some special applications, these constructions do not form the basis of or
dinary pneumatic tires. There are several reasons for this. It is necessary to
demount the tire from the rim to repair punctures or other accidental
damage. Manufacture of the normal demountable casing is far easier, bet
ter, and cheaper than any form of tubular tire. Specifically, a tubular tire
has to be made substantially in its final shape so that molding and consoli
dation of the inner surface and plies of the tire are difficult. These layers
are readily consolidated by the diaphragm or curing bag in the normal
casing of the demountable tire—a device which cannot be employed in the
manufacture of a tubular tire as it would be virtually impossible to remove
from the finished tire.
Tubeless tires, that is tires which are designed for use without a separate
(removable) inner tube, have a thicker inner lining inside and integral
with the tire casing—usually of a material of low permeability.
In order that the tire casing can be demountably fitted to the wheel rim,
the layers of the high modulus cords or filaments are turned around bead
coils made of a number of turns of high tensile, hard drawn steel wire, lo
cated at the inner edge of the tire sidewalks as shown in figure 4.5. Al
though in cycle tires a single wire bead coil can be used, as in figure 4.6a,
beads for pneumatic tires are usually comprised of many turns of hard
drawn steel wire. There are several reasons for this fact. The wire-drawing
process ensures that the tensile strength (100-150 tons/sq. in.) is several

(a)

(d)

FIGURE 4.5 Essentials of bead construction.


(a) Low turn-up construction.
(b) High turn-up construction.
(c) Overlap construction.
(d) Detail of a typical bead.
210 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.6 Bead wires.


la) Single bead wire for cycle tires.
(b) Single wire wound to produce a multi-turn coil.
(c) Tape of several wires wound in layers,
(d) Multi-wire woven tape.
(e) Cable bead.
1 1) Tape bead.

times that of the undrawn material. The multiple wire feature ensures a
degree of flexibility essential to tires fitted on a well base or drop-center
rim, and is a great help for ease of fitting and seating home on the rim base
even for tires fitted on detachable flange or divided wheels, as shown in
figure 4.7. The multiple wire bead coil can be made from an appropriate
number of turns of one length of wire as in figure 4.6b, or an appropriate
number of layers of a tape of several parallel wires embedded in hard rub
ber as in figure 4.6c, or an appropriate number of layers or turns of woven
tape, in which the several main bead wires are held together by an inter
lacing wave-wound binding wire shown in figure 4.6d. All of these bead
wire systems have the wires embedded in hard rubber, the wire having
been treated to achieve a bond to the hard rubber, usually by a copper
dipping process. The ends of the various wires in these multiple wire coils
are not joined in any direct way; the overall strength of the bead is
achieved by having an adequate overlap to ensure that the tension is built
up through the adherence to the surrounding hard rubber and con
sequently to the adjacent layer of wires as indicated in Figure 4.6c.
The single wire coil of a cycle tire is usually butt-welded to form a con
tinuous loop and often the welded joint is reinforced by a thin sleeve sol
dered to the wire as in figure 4.6a.
Another form of multiple wire bead coil is the cable bead made from a
single length of wire in which each layer of the winding is helically wound
at the opposite hand to the immediately previous layer shown in figure
STRUCTURE OF THE PNEUMATIC TIRE 211

FIGURE 4.7 Means of demounting pneumatic tires and extracting or inserting inner tube.
(a) Split or divided wheel.
(hi Well in the rim base or drop center rim.
(c) Demountable flange on the rim.
(d) Collapsible rim.

4.6c. This forms a more flexible type of bead than the other types of bead
coils and it also makes the ply turn-ups easier.
Since the tension developed in the coils of the upper layers of the lay
ered bead coils shown in figure 4.6e depends on the compression modulus
of the intermediate layers of hard rubber, these coils may be under lower
stress than the coils at the base of the bead. To minimize such effects when
particularly strong beads of low bulk are required, beads have been made
of hard drawn flat tape as in figure 4.61".
Other forms of bead have been used at times in the past but they are of
historic interest only, as they were either superseded by the wired-on type
or they failed to replace the wire bead.
The tensions in the casing cords set up by the inflation pressure are re
sisted by the tension developed in the steel wire bead coil. The material in
the bead which encases the steel wire coils is pressed against the rim flange
of the wheel upon which it is mounted by the inflation pressure within the
casing, some of this force being the axial component of the casing cord
tensions due to the direction of the sidewall at the bead as shown in figure
4.8. This reaction between rim flange and tire bead enables traction and
212 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.8 Forces pressing bead against rim flange on the wheel to obtain driving and brak
ing reactions.

braking forces to be transmitted by friction between tire bead and rim


flange. The magnitude of available traction or braking force is augmented
by the radial pressure between tire bead base and wheel rim seat due to
the load on the tire and also by the use of an interference fit between bead
and rim dimensions.
The height to which the plies are taken after they have passed round the
bead coil can be low as in figure 4.5a, or high as in figure 4.5b, or the plies
can extend across the crown as in figure 4.5c. In this case of crown overlap
the plies can be arranged to give extra casing strength in the crown region
compared with the sidewalls.
The choice between these alternatives rests on experience of fatigue life
of tires in the field and to some extent on manufacturing considerations. In ,
all cases where a number of plies finish in relatively close proximity the
endings are staggered as in figure 4.5d, or a variant of this, so as to cause a
gradation of the stress concentration and stress transfer to the main part of
the plies. The integrity of the tire structure depends on the security of at
tachment of these ply ends to the main part of the plies, which form the ^
tire casing, and this depends on the bond between the high modulus cords
or filaments and the low modulus material in which the filaments are em
bedded, and the bond between the various layers of low modulus materi
als in that region of the tire.
The steel bead coils may have a canvas wrapping. The bead almost al
ways has a form of packing above it to give a suitable path of return for
the edges of the plies to the main part of the carcass, as in figure 4.9a. If
this packing, or filler, or nipper as it is variously called, were absent the
tension would not build up smoothly from zero at the end of the cords to
the full value at the bottom edge of the plies of the inner part of the side-
wall. If the plies were wrapped round the bead coil only without the filler
being present, as in figure 4.9b, only the cranked portion, that is the por
tion wrapped round the bead coil, would be of use in resisting the cord
tensions, and the construction might as well be as in figure 4.9c. In this
case the only way that the cord tensions would be resisted and the move
ment of the plies prevented would be by the stresses set up in the bead coil
acting as a ring with a torsional moment applied along its periphery. The
resisting moment and the torsional rigidity of even a solid ring, when
loaded in this manner, is not great and consequently the usual multiple
STRUCTURE OF THE PNEUMATIC TIRE 213

FIGURE 4.9 Conventional layout to give filler with good return transition for ply end.
(a) Satisfactory design with filler.
(b) Unsatisfactory design with no filler.
lc) The effectively stressed part of design (b).

wire bead coil would be quite inadequate to resist the ply tension. Hence
plies are turned up round the bead and passed across the bead filler to ad
here to the main plies as in figure 4.9a.
In tires which have a large number of plies, two or even three bead coils
are employed, with each bead usually having about the same number of
plies around it as in figure 4.10. Multiple bead coils are not used in tires
with less than eight to twelve plies and so they appear only in tires for
trucks, earthmovers and aircraft.

FIGURE 4.10 Multiple bead coils for tires with a large number ofplies.
214 MECHANICS OF PNEUMATIC TIRES

The bead region commonly has a layer of textile material, such as wo


ven fabric, wrapped around it before the outer rubber covering is applied.
This is shown in figure 4.5d. The function of this chafer, as it is called, is to
minimize the movements of the tire bead surface in contact with the steel
rim flange in the region in which the tire cords are caused to move within
the tire wall due to the application of the load on the road. In the area
where the tire bulges due to load there are radial and longitudinal move
ments of the tire wall which are constrained by the chafer, and so abrasive
wear is eliminated at this point. In some tire designs rubber chafers of ap
propriate hardness are suitable for this function.
In a large proportion of tire designs, security of the attachment of tire to
rim is augmented by an interference between bead and rim diameters.
This interference is always employed on tires fitted to one piece rims. It is
essential for tubeless tires, to ensure an airtight seal between bead and rim,
and even with tube tires it is more necessary with modern wide base rims
than it was in some of the earlier narrow rim widths, because of the rela
tive magnitudes of the axial component of casing tension. The interference
causes compression in the material under the steel bead coils, and this in
turn sets up a tension in the coils which adds to the effects of inflation.
Sometimes dimensions are so chosen that the tension in the bead coils due
to interference fit is greater than that due to casing inflation tensions. In a
large proportion of designs for low pressure tires, such as passenger car
tires, the interference fit is so chosen that the tire has to be inflated to a
pressure substantially higher than its operating pressure in order to elastic-
ally stretch the coils sufficiently to permit the bead to move up the seat.
This ensures that the beads seat home against the rim flange during the fit
ting of the tire on the rim.
The axially outward component of the casing tensions press the bead
against the rim flanges and set up a tension in the rim base, shown in fig
ure 4.1 la. In the little-used tubular tire design this tension would be taken
either wholly or in part by the casing cords on the inner periphery of the
toroid, as illustrated in figure 4. 1 1 b. However, it should be noted that
tubular tire designs require anchorage points to the wheel, spaced apart as
in the manner of the beads of the conventional form of tire. Without such
anchorage points the tubular tire would roll sideways and so provide only
low resisting forces to sideway drift on the road. It would in consequence
have only poor guiding properties. The apparent digression to discuss the
almost unused tubular tire design serves to draw attention to important as
pects of the function of the bead and of the high modulus wire coils in the
bead of the conventional tire.
Demounting of a conventional tire is made possible by either the provi
sion of a well in the rim base, also known as drop center, or by the use of a
detachable flange, or by having the wheel in a divided or split form. These
are shown in figure 4.7.
The shape of the tire bead and rim on to which it fits can be divided into
two main categories, the 5° and 15° taper rims shown in figure 4.12. The
15° taper rim design is the more recent design, having been introduced in
medium size truck tires in the United States about 20 years ago. Variants
of the more generally used 5° taper rim have existed in the past, since it
developed from a cylindrical base, or 0° taper. There have been one or
two examples in the truck tire range with a taper angle of 8° to 13°. Al
though such rims may still exist in service, the tires are designed for
mounting on the 5° taper.
STRUCTURE OF THE PNEUMATIC TIRE

FIGURE 4. 1 1 Axial tension in rim base.


(a) Tire and wheel.
(hi Tubular lire.

The flange height of the rim is higher for the 5° rim than for the 15°,
and the tire beads have some necessary differing disposition of internal
components. Tire sectional heights are shallower for the 15° rim than the
5°. For example, the 22.5 in. rim shown in figure 4.12 uses a tire of the
same outer diameter as the 20 in. rim also shown there.
All tires in regular use have two equal size bead coils but there is no the
oretical reason why the two bead coil diameters should not be different.
One could produce technical arguments in favor of such designs, and in
fact various tire designs of this type have been the subject of patents but
none have gotten beyond the experimental stage. One practical objection
against them, and an objection raised against any asymmetrical or direc
tional tire feature, is that they have to be fitted one specific way. If the un
equal rim diameter is coupled with a directional tread pattern feature then
two different moulds and tire types are required for any given vehicle.
The technical arguments which have been put forward for the use of a
smaller diameter bead seat on the side of the wheel away from the vehicle,
and a larger diameter bead seat on the inner side are the lower risk of curb

FIGURE 4. 1 2 5° and 15° taper rim flanges drawn in similar relation to the wheel axis of
revolution.
216 MECHANICS OF PNEUMATIC TIRES

damage with a long sidewall on the outside, and greater diameter for the
brake on the inside of the wheel. Further arguments stem from the fact
that during cornering of a vehicle the wheels on the outside of the curved
path, which are the ones which carry the greater load, are usually caused
to camber by the suspension link work and this camber is in the direction
to increase the deflection of the outer sidewall of the tire, and hence to
lessen the deflection of the inner sidewall, the terms inner and outer being
in relation to the vehicle here. The fact that the tires on the inside of the
curve are usually cambered so as to shorten the short inner wall more than
the long outer wall of the same tire is not important, it is argued, since the
load on this tire is lower and so the percentage deflection of the short inner
wall of the tire on the inside of the curve need not be any greater than the
percentage deflection of the longer outer wall of the tire on the outside of
the curve. Even so, no such tires have gotten beyond experimental investi
gation.
So far, tire casing and beads have been discussed in a general descrip
tive way. The third important component of a tire is the tread.
The tread performs several functions. It is the only part of the tire to
come into contact with the road surface. It provides a wear-resistant layer
and also protects the casing. It provides frictional contact with the road
sufficient to transmit driving, braking, and cornering forces. These fric
tional forces may reach a value equal to the load carried by the tire. The
tire tread carries a pattern of such character and detail design as to ensure
adequate removal of water and other contaminants from the road surface,
so as to maintain an adequate level of frictional adhesion between tire and
road over a wide range of operational conditions.
For tires intended for operation on soft ground, the character of the de
formation of the ground and the laws of soil mechanics determine the
form of the tread pattern.
The only type of material which has been successfully used as tread ma
terial is rubber or a rubber-like material, that is, a long chain molecular
material or polymer of a modulus comparable with that of the matrix in
which the filaments are embedded in the rest of the structure. The tread
polymer has to be reinforced with suitable ingredients such as carbon
black to obtain the required abrasion resistance but this is a subject in it
self and will not be discussed here because detail changes in this material
do not significantly alter the tire structure required. Experience in the de
velopment of spring wheel devices shows that rubber, natural or synthetic,
has no competitor as a tread material.
These materials automatically give good friction on dry roads and some
aid the achievement of good friction on wet roads. This will not be dis
cussed here since it is covered elsewhere in this book.

4.3. Flexible Filament and Soft Matrix Constructions


Although the patent literature contains a wide variety of alternatives to
the use of filaments or cords in tires, all commercially successful tires are
now built as a series of layers of flexible high modulus cords encased in a
low modulus rubber or rubber-like material, the cords in each layer being
in a given path or direction and substantially equi-spaced and parallel.
A conventional cross bias casing or carcass is built of two or more, usu
ally an even number, and most commonly four, layers of parallel cord fila
STRUCTURE OF THE PNEUMATIC TIRE 217

men ts, the direction of the cords in a layer being at an angle to the princi
pal axes of the tire. Half of the layers have the cords at a positive angle
and half at a negative angle as shown in figure 4.13. Usually alternate lay
ers have cords at the opposite hand, thus giving the cross bias effect be
tween adjacent layers at all points in the tire as in figure 4.14a. However,
tires have been made with adjacent layers of cords parallel. In the case of
four-ply tires the middle two plies are parallel, and in the case of more
than four plies, the ply arrangement is, say, AABBAABB etc, or, AB
BAABBA. Some of these are shown in figures 4.14b and 4. 14c. Tires have
been made with an odd number of plies, and even tires with three plies
have been commercially tried but they are the exception rather than the
rule.
Usually the crown angle of the cords in the two directions are equal in
magnitude. The crown angle is the angle between the path of the cord and
the line along the tire periphery defined by the intersection of a plane at
right angles to the axis of rotation of the tire and the highest point of the
tire cross section, as shown in figure 4.15.
Tires of cross bias construction are rarely made with a crown angle ex
ceeding 38°. Lower angles are used, particularly when high speed or other
special performance characteristics are to be achieved. Tires intended for
vehicles on public roads are rarely less than 10° lower than the above fig
ure, although tires with crown angles of the order of 20° lower may be
used for very high speed vehicles restricted to use on tracks.
Cross bias tires with a crown angle substantially higher than 38°, al
though capable of giving a good comfort, are not used because of their low
cornering power and consequent poor guiding characteristics.
In order to give increased protection to the casing under the tread and/
or to increase the casing strength in the crown region of a cross bias tire,
one or two layers of cords are sometimes incorporated substantially paral
lel to the cords in the other plies but extending only approximately the

FIGURE 4.13 Cross-bias tire.


218 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.14 Cord ply arrangements.


(a) ARAB- - the usual.
(b) AABBAA.
(c) ABBA.
(b and c are called parallel ply construction', i

width of the tire tread. This is shown in figure 4.16. Such layers are called
breakers. Sometimes the cord spacing in the breaker plies is greater than
in the main plies, sometimes the spacing between layers is greater than be
tween the main plies, and sometimes insulations of differing modulus are
incorporated above, between, or below the breaker cords. These variants

FIGURE 4.15 Cord crown angle.


STRUCTURE OF THE PNEUMATIC TIRE 219

FIGURE 4.16 Breaker layers in cross-bias lire.


The cords in the breaker layers are the same crown angle as the casing cords.

are decided from service experience and relate to the fatigue life rather
than the mechanical behavior of the tire as a structure. A tire of cross bias
casing construction with a breaker of the same angle as the casing is not
detectably different in mechanical behavior from a tire without a breaker
but otherwise of similar construction, i.e., cord angle, number of plies, etc.
The above comments relate to the conventional cross bias tire, which
has been used from the beginning of the automobile era. However, if the
cord angles in the breaker layers are substantially different from those in
the main plies, and the breaker is made of either higher modulus cords
than the casing or of more layers than the casing, then the breaker con
struction has an important bearing on the mechanical properties of the
tire. Tires of this type are known as bias belted tires (sometimes incor
rectly called semi-radial tires) and have been popular mainly in North
America during the last 10 years; these tend to display properties more
akin to the cross-bias tire than the radial tire now to be discussed.
For about 25 years in Europe, and more recently in North America,
tires of a construction very different from the cross bias construction have
been produced and successfully used in large quantities in many cases
completely displacing the cross-bias tire. In these newer tires the cords or
filaments in the casing are disposed in a radial, or substantially radial, di
rection giving a 90° bias or crown angle in relation to the axis of rotation
of the tire. They also use a breaker or belt of several plies of cords fitted on
top of the casing under the tread, and laid at various crown angles, two of
the layers at least having a low crown angle of the order of 20° as shown
in figure 4.17. These tires are commonly called radial tires although a
more correct description is rigid breaker, radial ply tires. An alternative
name is belted radial ply tires. The rigid breaker or belt is essential to the
220 MECHANICS OF PNEUMATIC TIRES

FIGURE 4. 17 Radial ply rigid breaker tire.

functioning of the tire. Without it a radial ply casing can become unstable.
When inflated to a high pressure, determined by the magnitude of the
slight irregularities in cord spacing, differing extension of the rubber be
tween the cords at different points around the tire permits the cords to
move out of a radial plane and the tire periphery develops into a series of
severe buckles. Such an instability is to be expected if the comparison is
made with a tall pile of interspersed rubber and metal washers under com-
pressive load.
The concept of a radial casing with a reinforcing band or belt on it in
the crown region, although developed successfully only recently, is much
older since a British patent dated 1913 describes a tire in which the rein
forcing belt is placed under the casing cords. These early designs were
never brought to a successful development state.
The objective of the belt or breaker of low cord angle is to provide rigid
ity to the tread against the distortions in the lateral direction which are set
up during cornering, and so to reduce tread wear under this condition of
use and to improve the handling.
The conflict between a nondevelopable surface, such as a surface of rev
olution of curved cross section, which in consequence is a surface of
double curvature, and a flat plane surface is an important factor in the de
cisions of designers to employ cords embedded in rubber for the tire struc
ture.
The simplest tire-like form which can be expressed in simple geometry
is a toroid. If the curved cap is flattened from its original shape to a flat
plane without altering any other part of the tire periphery, as shown in fig
STRUCTURE OF THE PNEUMATIC TIRE 221

FIGURE 4.18 Toroid in contact with flat plane.

ure 4.18, then the length along the surface in the direction of the longer
axis of the patch, originally aD, is compressed to a length 2a. Now
a2 = S(D - S)
and
2a
a = sin

Noting that

6 40
and by putting
la
x=
D
we have
Original length = •:—5L:—:—^-5—
length along arc = -^—
aD . ,
—rf—:-^— 1, + Longitudinal
, Strain
New length length along flat 2a

, , 2 d(D-S) 6 8\D-dY
3 IT S D*
x 1 + 28
3D \+ -%-
3»s
222 MECHANICS OF PNEUMATIC TIRES

Consideration of the dimensions across the patch leads to


Off T£
^r- ~ 1 + TIT = l + Transverse Strain.
2.D jfi
It follows that the surface may undergo compression in both directions at
all points in the area under discussion while simultaneously the size of the
elliptical contact area of this patch is unaltered. It also follows that if the
cap were flattened without compression in either of its principal axes, the
outer edges of the elliptical patch must extend or tear.
In the case just discussed it was assumed that the pan of the tire outside
the contact patch is undisturbed. This might be the basis of a challenge
that the example, as discussed, does not establish the point being made. It
is not, as can be seen by considering a sphere. If the structure under con
sideration were a sphere made of isotropic material, instead of a toroid,
the contact patch would be circular, not elliptical. The actual circular con
tact patch (ignoring transition curves due to shell bending stiffness, which
in an appropriate case can be considered to be a second order effect) will
lie between that given by the intersection of the sphere and the "contact
plane" and a circle of diameter

-I
times as large, as shown in figure 4.19. The former corresponds to a hypo
thetical case in which all the distortion is in the cap or part of the shell
within the contact patch, and there is no tension along the periphery of the
contact patch. The latter case corresponds to the hypothetical case in
which there is no radial compression in that part of the shell forming the
cap or contact patch, but in this case the periphery of the patch is ex
tended in the ratio (1 + 2δ/3S). This implies that the part of the sphere
outside the contact patch is strained.

FIGURE 4.19 Sphere in contact with flat plane.


Actual contact patch radius will be between the values ab and cd. where d is the intersection point of tphere and plane
and ab equals arc of ad
STRUCTURE OF THE PNEUMATIC TIRE 223
b' b"

d"
FIGURE 4.20 Two-ply high modulusfilament embedded in low modulus matrix under biaxial
extension.

In the actual case the extensions will lie between zero


and 2δ/S, and at least one of the strains will be of that order if another is
near zero. If they are all of similar magnitude they will be of the order of
δ/3S. If the sphere diameter S is similar in magnitude to the diameter of
tnTcross section of the toroid discussed above, and δ/S or δ/H is taken to
be of a magnitude of acceptable tire operating deflection, i.e., 0.15 to 0.35,
then the resulting strains are greater than any crystalline metallic material
can withstand within its elastic range or fatigue limit. Clearly if the sheet
or surface were isotropic it would have to be made of a material having a
high extensibility before fracture. This demands a low modulus, and such
a structure would change dimension on inflation rather like a toy balloon.
The alternative possibility of a material with a medium to high modulus
would require a very high breaking strength and a very high yield point,
and no such material has been found. It is for this reason that proposals
for thin shelljoroidal spring wheels of metal or plastic have all so far been
abortive.
—fiToTaer to get a rough idea of the deformation of a shell structure of
double curvature made of high modulus filaments embedded in a matrix
of low modulus material, consider the following case chosen for ease but
reasonably accurate discussion. Figure 4.20 represents a two-ply sheet,
made of one ply of filaments parallel to two parallel sides of the rhombus
A BCD, and the other ply with filaments parallel to the other pair of sides.
Imagine that this sheet of composite material undergoes a biaxial exten
sion of the order of magnitude which would arise at the principal axes of a
toroidal indentor stretching the originally flat sheet to the shape of the cap
of double curvature discussed above. Consider two filaments AB and DC
in one layer or ply, and two filaments BC and DA in the other ply, before
stretch. If one assumes that the filaments are inextensible, then after exten
sion the filaments move to a′b′ and <?c' in one ply and b"c" and d"a" in
the other ply. These movements are permitted by the low modulus matrix,
which develops shears as indicated by the tilt of the lines a'a", b'b", c'c",
and dd" from their original A'A", B'B", C'C", D'D" as shown in figure
4.21. Within a ply there is shear and extension indicated by comparisons
of a'b' and A'B' etc.
224 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.21 Perspective view offigure 4.20 to show interply shears.

Change of cord length, assumed not to occur in this example, would of


course exist in many practical cases. This would cause some modification
of the strains in the matrix but in detail only.
As the local flattening of a surface of double curvature involves a proc
ess the reverse of the above, the problem is analogous but results in the
initial tensions being reduced during the process instead of being in
creased. In some cases the cords will buckle in the process but because of
their close juxtaposition within a layer, and because of the layered con
struction, the buckles are of a controlled character. The cords are elastic-
ally supported by the matrix in relation to the neighboring cords, and the
wavelength of the buckles is determined by the consideration of a cord as
an elastic beam on an elastic foundation.
This arbitrary example is artificial because it has been considered in the
flat plane only, for ease of discussion and presentation. However, it does
show how plies of high modulus cords or filaments embedded in a low
modulus matrix can permit distortion in a structure which would not be
possible if the structure were formed by an isotropic sheet of that material.
It is clear that in changing from a surface of double curvature to a plane
surface the high modulus filaments must bend. They must also be flexible
enough to be turned round the bead coils during the manufacturing proc
ess. Furthermore, during tire usage, conditions arise where flexibility or
the ability to bend is vital. The most severe bending would probably be
during accidental running of a tire in a deflated state. Reasons such as
these require that the high modulus filaments must be flexible high modu
lus filaments.
Flexibility of the high modulus filaments is obtained by restricting the
diameter of the filament or/and using a multiple fiber or multiple strand
cord in which each separate strand of the high modulus material is suffi
ciently thin to keep the stresses below both the permanent set and the fa
tigue limit under any condition which might be met—even if the occasion
were infrequent. Typical cord cross sections are shown in chapter 2 of this
book.
STRUCTURE OF THE PNEUMATIC TIRE 225

It is suggested here that a general requirement might be expressed in the


following form. The high modulus filaments shall have, by the choice of
strand or fiber diameter and construction or cabling, sufficient flexibility
that they are not overstressed as judged by any appropriate criterion, in
cluding fatigue and permanent set, if bent as much as it is possible to bend
the low modulus matrix of the size and character into which they are em
bedded. It follows that small tires with thin walls need the finest strand fil
aments in cabled cords, and that coarser strands may be permissible in the
thicker walls of large tires.

4.4. Tire Construction Methods


The conflicting requirements of limited expansion of the structure dur
ing inflation, coupled with the need for internal shear movements within
the sheet forming the surface or envelope when a tire structure of double
curvature is flattened at the ground contact patch, necessitate the use of a
multiplicity of relatively inextensible, but flexible filaments embedded in
an easily distortable matrix. Accepting this, it can be seen that, in theory,
there are a number of alternative ways of building a tire.
All methods of making a tire of these composite materials involve a
building process followed by molding and vulcanization. The building
process is the stage where the materials or components are placed in the
required appropriate relative position one to the other, and will be dis
cussed further below.
The molding process is needed for the compaction and consolidation of
the various components, for example embedding the higher modulus fila
ments in the low modulus matrix to eliminate voids and to obtain the in
timate contact necessary to secure adhesion between filaments and matrix,
and also between layers of the matrix itself. The vulcanization process is
required for changing the state of the rubber or rubber-like polymer used
for the low modulus matrix, from the uncured plastic state necessary for
the building and molding stages of the manufacturing processes to the
elastic state essential for effective tire performance in the final product,
and for the formation of the adhesive bonds between filaments and matrix
and between all layers of matrix within the tire. The uncured plastic state
of the rubber-like polymer has a feature of value during the building stage
and also the molding stage, that of "self-tack" whereby two contacting
surfaces adhere and, even if separated, readhere on recontact.
It should perhaps be mentioned at this point that although a fairly wide
range of materials have been successfully used as the high modulus fila
ments e.g. cotton, rayon, nylon, polyester, glass, and steel, which range
from organic, long chain polymeric to crystalline materials, the low modu
lus matrix has always been a long chain polymer of rubber or rubber-like
characteristics. It is this material which requires vulcanization. Although
many of the materials used in the filaments are affected by the heat used in
vulcanization, not all are, and in practically all cases the vulcanization
stage is not required to bring the cord material into its required final phys
ical state.
The molding process also fulfills at least one other function, namely the
formation by molding of the tire tread pattern. In many of the more mod
ern methods of building, the shaping of the tire from the assembled shape
to the required final shape is carried out during the molding process.
226 MECHANICS OF PNEUMATIC TIRES

The alternative ways of building a tire casing can be enumerated as fol


lows:
A. The tire could be manufactured on a former or pattern substantially
of the shape of the desired finished shape of the tire, as shown in figure
4.22a. Let this be denoted process A.
B. The tire could be made on a former of a shape other than that of the
desired finished tire shape. The former could be a simple cylinder, as
shown in figures 4.22b and 4.22b1. In this case the change of shape can be
effected as a separate process prior to the molding stage, called process B1,
or during the molding stage, called process B2, using the same molding
equipment in which the vulcanization is carried out.
C. Part of the tire could be manufactured on a former different from
that of the final shape, and that part of the tire shaped to substantially that
of the final tire shape, and then the remainder of the tire components
could be added. Let this be called process C.
In each of the above cases the tire casing can be built in one of the follow
ing alternative ways, or some components built in one way and some in
another.
a. The tire casing could be assembled from sheets of high modulus fila
ments or cords embedded in a low modulus matrix, the sheets being
produced by a separate previous process which ensures that the fila
ments are straight, parallel and equispaced, within the rubber. These
sheets can be laid on the building former shown in figure 4.23a ei
ther
(1) individually, as shown in figure 4.23a1,
(2) in pairs having been previously put together so that their angles
are complementary, as shown in figure 4.23a2,
(3) in pairs having been previously put together so that their angles
are asymmetric,
(4) in four or more layers having previously been assembled so
that alternate layers are at complementary angles, or
(5) in four or more layers having been previously assembled with
their angles in some relationship other than alternately com
plementary.

FIGURE 4.22 Manufacturing processes I.


(a) Casing built on "core" former substantially the size and shape of the finished tire.
(b) Casing built on a "low crown" former—a different shape from the finished tire, and distorted to the required shape in
i subsequent process.
(hi) Casing built on a Mat or cylindrical former and distorted to the required shape in a subsequent process.
STRUCTURE OF THE PNEUMATIC TIRE 227

FIGURE 4.23 Manufacturing processes II—tire casing building methods.


(a) Plies of material with multiple cords cut to dimensions and wrapped around a former,
•i, single ply.
a;, pocket or paired plies at opposite bias angles.
(b) Plies with cranked or curved cord path.
(c) No illustration (see text).
(d) Single cord wound on former,
dj, as a simple winding.
ds, as a wave winding.
(e) No illustration but as (d) using narrow strip of several cords.
(f) Strip of material wound on former.
(B) Band of pre-rubbcred cords wound around and through the bead coils held apart,
(h) Same, but for single cord,
(i) Peg winding—it cords at one bias.
12, cords al croa bias.
228 MECHANICS OF PNEUMATIC TIRES
b. The rubbered multifilament sheets used for the assemblies (listed
above in a.) could be distorted so that the cord paths were curved
from the straight, as shown in figures 4.23b.
(1) Before assembling into pairs.
(2) After assembling into pairs.
c. The rubbered multifilament sheets could be used as in a, but the
curved effect or an asymmetric angle effect could be gotten by a rela
tive rotation of one tire bead in relation to the other during building,
shaping or molding.
d. The cord or filament could be wound as a single rubbered cord on a
former appropriate to building method A or B,
( 1 ) as a simple uniform winding.
(2) Great circle winding, with a slight advance of location at each
turn, resulting in spaced winding, as in figure 4.23d2, or
(3) controlled wave winding, as in figure 4.23d3.
In all cases of d the cord can be rubbered at the building machine by
passing it through an extruder die.
e. A narrow band of pre-rubbered continuous cords can be wound on
the former as a strip, with all of the possible variations listed in d
above.
f. A band of two or more layers of cords can be wound on the former
as shown in figure 4.23f:
(1) with the constant angles differing at different places across the
former.
(2) with differing numbers of layers at various places.
(3) with all the possible variations in d listed above.
g. A band of pre-rubbered cords may be wound around and through
the bead coils suitably held apart, the bobbin carrying the band of
material passing through the bead coils during the winding process.
This is shown in figure 4.23g.
h. As in g but using a single cord rubbered at the building machine.
i. A single cord wound from bead to bead, the cord being located by a
series of pegs or slots:
STRUCTURE OF THE PNEUMATIC TIRE 229

(1) the cords all being parallel to one another in each layer as
shown in figure 4.23i1,
(2) the cord progressing forward at the opposite bias at each peg,
as shown in figure 4.23i2.
j. As in i but using a narrow band of several cords.
The above classification seems to cover all possible ways of building a
tire. All of these cases and combinations of cases seem to have been the
subjects of patent applications. Many have been tried and used for some
tire application or another at various times.
The vast majority of passenger car and truck tires were originally made
using processes A and a2, but quite early in the history of the industry the
process became Bl and a'2, and quite soon afterwards B2 and a2. Today
the vast majority of car tires, truck tires, earthmover and aero tires, and
some belted cross bias casing tires are made this way. All production
radial ply belted tires are made by processes C and a1, a2 or a3.
Most of the other methods have not gotten beyond the experimental
stage, and practically all lack the flexibility of the current standard meth
ods. One method which has become standard, that of g and B2 has been
used extensively for the manufacture of cycle tires. It is restricted, by its
character, to tires whose section width is much smaller than the bead di
ameter. The bobbin carrying the total material for the tire must be capable
of passing through the bead coils while they are held apart and tensioned
by either four drums or a system of arcuate guides, so that the portions of
the bead coils between are straight and sufficiently tensioned so that the
winding of the rubbered cord does not cause the beads to sag towards each
other. There is another restriction of this method however, since the choice
of bias angle is linked with the strip width being wound and the bead coil
periphery. Furthermore, the bead coils must be sufficiently flexible for the
passage around the roller system. The problems specific to this method
have been mentioned as an example of the fact that most of the alternative
methods set out above have restrictive characteristics peculiar to each one.
The shaping and molding stages are illustrated in figure 4.24. There are
two main alternatives. The curing bag or bladder is inserted into the al
ready shaped tire, as in figure 4.24a1, or into the unshaped tire as in figure
4.24a2, during shaping by a machine process. Afterwards the tire is vulcan
ized in a mold in a press.
The more modern process is one in which the unshaped tire is in
troduced into a special press in which a cylindrical diaphragm shapes the
tire as the press, carrying the two half molds, closes as illustrated in figures
4.24H, and 4.24b2. The inflated diaphragm consolidates the tire during the
molding and vulcanization process in the heated mold.

4.5 Cord Path—Practical Factors Determining Selection


Although in theory any cord path is possible in any molded shape there
are several factors which restrict the range of practicable cord paths.
At the outset a choice must be made as to whether the cords are to be
material, or (II) individually, or (III) in a group relatively few in number
to produce a band or tape of rubbered cord.
In the first case (I), sheets of several feet in width and hundreds of yards
230 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.24 Manufacturing processes III—Molding and vulcanization.


(a) Shaping and fitting curing bag on the bladder as separate operation.
(b) Shaping as part of the diaphragm molding and vulcanization process.

in length, comprising a large number, say one or two thousand, cords


across the width are produced in a continuous calendering process sepa
rate from the tire building. Sheets, called plies, of appropriate dimensions
are bias cut from the calendered sheet so as to give the required angle of
cord direction to the cut edge. This is the most common method and is a
very flexible process, since a wide variety of ply dimensions can be pro
duced as required. The only real limitation is that the initial rubbering
process must be made with the cord spacing at its required amount, and
the correct thickness of rubber applied.
In the other two cases (II) and (III), the cord would be rubbered as it is
wound on a former of appropriate shape at the tire building machine.
STRUCTURE OF THE PNEUMATIC TIRE 231

The factors which control the speed of covering the high modulus cords
with low modulus rubber-like compound are the temperature and temper
ature-time cycle and the vulcanization and scorch characteristics of the
rubber-like material, along with the characteristics of the bonding process
of rubber to cord. These place an upper limit to the speed at which the
cord can pass through the rubbering stage. Clearly, without going into sec
ondary detail consequent on the differences in the various processes in
volved, rubbering a thousand cords simultaneously is more expeditious
than rubbering one or a score of cords at once. It is for this reason that the
processes using calendered sheet material dominate in commercial tire
production.
Processes involving curved cords, or other paths than straight, in the ini
tial flat sheet even if produced initially straight and subsequently cranked
in a special process, raise the problem that limitations on the closeness of
spacing in the low angle cord region demands relatively wide spacing in
the high angle cord region. The geometric necessity for this is clearly seen
in figure 4.25.

4.6. Calculation of Cord Length


The integrals for finding the length of a line drawn on a solid of revolu
tion are known explicitly only in very special cases, such as the constant
angle helix on a cylinder, or a great circle on the surface of a sphere.
In the case of a tire, even if the tire cross section were circular, which it
never is, the integration does not lead to a solution in a closed form. In the
general case of a nonci rcular cross section the integral cannot be written
explicitly, let alone evaluated and its solution stated.
On the other hand, if the path of a cord is traversed step by step, and the
length and direction of each of the steps is recorded, it is possible, pro
vided the cross-sectional shape of the tire is known, in numerical, graph
ical or mathematical terms to resolve the progress along the path into any
definable coordinate system. This can be done whatever the cord path and
whatever the tire cross-sectional shape.
This resolution into a known coordinate system is analogous to naviga
tion at sea, with the difference that the map is not a plane, or a sphere, or
generalized three-dimensional space. It is the surface of the solid of revo
lution of a cross section defined by the tire shape specific to the case in
hand. Whatever the path or whatever the cross section, the progress in the

FIGURE 4.25 Cranked cord limitation on cord spacing.


The number or cord ends per inch in the high ingle region ii limited by the limitation of closeness of cord spacing in the
low angle region.
232 MECHANICS OF PNEUMATIC TIRES

FIGURE 4.26 Definitions of radius R and cord angle d>.

desired coordinate system is recorded by resolution of the known step in


the known direction, followed by simple summation in each coordinate.
Integrals expressed as mathematical functions do not arise. In effect, the
simple summations of the coordinates correspond to, and replace the need
for, the integration along the path itself.
There are two simple relations between /?, the radius, and cos <j>, both as
defined in figure 4.26, which hold over all possible cord paths in a tire. For
this reason a graph of R vs. cos <> is an important diagram. Note here that
Φ is defined as the angle between the cord tangent and a parallel on the
tire surface, while R is the radius to the axis of rotational symmetry. The
chart obtained by plotting R versus cos <£, or a formula relating the two
quantities, or a numerical tabulation of pairs of values, is the fuU descrip
tion of a cord path or as appropriate, of the movement of a point on the
cord during processing. Pantographing movements and geodetic move
ments are linear and hyperbolic relations respectively on this chart. On the
other hand, the geometry of the tire surface is completely described in
terms of R versus X, where X is the distance along the meridian, that is the
distance along the surface measured from the tire crown in the plane of a
cross section of the tire as shown also in figure 4.26.
Given the cord path by means of R versus cos <j> on a former or mold of
known R vs. X, then if the tire is shaped by a process involving movements
expressed in terms of an appropriate transformation of R vs. cos <j>, it is im
mediately possible to determine the R vs. cos <f> of the cord path in the fin
ished tire.
It is also known that the R vs. cos <> chart can be used to diagnose the
character of the cord path in a tire, or to elucidate the character of the
processes occurring during the shaping and molding stages of tire manu
facture.
Consider any surface of revolution on which any cord path is drawn.
Consider a tire made of two layers, one layer comprising a large number
of cords of the same cord path and these cords being equispaced and en
compassing the whole surface of this layer, and the second layer of the
same system of cords but of opposite hand. Furthermore, consider that the
cords in these two layers are pin-jointed one to the other at their crossing
points. It has been shown in reference [1]2 that the shape of this network
2 Figure in bracket! indicate literature references u the end of this chapter.
STRUCTURE OF THE PNEUMATIC TIRE 233
Trellising or pontographmg

Crown- Final Shape


of Cross- bios Tire

/\\Range of Values
for a Geodetic
Crown- Final \T'r«
Shape of Range of Values
Geodetic Tire for a Cross- bias Tire

Hyperbola- Geodetic -

COS OS

FIGURE 4.27 R-cos <> chart.


Plots for conventional and geodetic paths.

can be changed, and provided that the surface is maintained to be a sur


face of revolution around the original axis, the treUissing or pantographing
of the cord structure will ensure that all times
COS C)
= c,
as shown in figure 4.27, where C, is a constant.
If the solid of revolution is a cylinder and if the helix angle is constant,
R and <;> are the same at all points on the path, so that one point on the R
vs. cos <f> graph represents the whole helical path. This corresponds to the
path on a cylindrical building drum. If the tire casing is shaped up by
pantographing without slippage or extension of the cord then the cord
path will be described by a portion of the straight line through this initial
point and the origin, as illustrated also in figure 4.27.
It follows from a study of geodesies on any surface of revolution that the
relation between R and cos <> for points on such a path, which always fol
lows the shortest route between any two points on the surface, is
R cos </> = C2
where C2 is a constant.
The fact that both of these relations may be displayed simultaneously
on the R vs. cos »j> chart makes such a chart important. The plot on the R
vs. cos tj> chart is independent of the cross-sectional shape, since a curve
obtained by treUissing or pantographing always lies along a line through
the origin, while a geodesic path always appears as a hyperbola asymp
totic to the coordinates of the chart. This is shown in figure 4.27. This per
haps is made clearer if Log R is plotted against Log cos <j>, as in figure 4.28.
Here, the two lines are straight at 45° to the coordinates, and are ortho
gonal to one another.
If in the pantographing system the cords extend, then the R vs. cos <j>
values which are obtained lie on a new straight line through the origin
234 MECHANICS OF PNEUMATIC TIRES

SLOPE -H

TRELLIS OR
PANTOGRAPHING

LOG COS <J>

FIGURE 4.28 Log R-log cos $ chart.

with a slope of (1 + cord strain) times the slope of the line describing the
original unstretched cord path, so that R = R(1 + ∊) at a given value of cos
<j>. This is shown in figure 4.29.
The above discussion relates to the characteristics of an actual path of a
cord on a surface of revolution. The presentation of information on a set
of coordinates of R and cos <f> has another usage, that of depicting the
changes or movement during shaping of the tire casing, or other manufac
turing processes. Movement of the point on the R vs. cos <f> graph, repre
senting a specific point on the actual tire cord, can be examined if the fol
lowing facts are noted. As the cord lattice pantographs, the point on the R
vs. cos <j> chart moves along a straight line through the origin, as illustrated
in figure 4.30. As a cord stretches, the point on the chart moves along the
hyperbola passing through the point, as in figure 4.30b. If slippage occurs
the angle remains constant but the radius R changes. This results in a ver
tical displacement of the point, as in figure 4.30c. For further discussion of
this concept see reference [1].

b-R vs COS 0 with cord Strom


0-R vs cos * without cord strain J

FIGURE 4.29 R-cos <fr Effect of cord stretch.


If Uw cord length chuifes in the ratio (I + c) the R-coi * line changes from a lo * where the slope of* il (I + c) I
STRUCTURE OF THE PNEUMATIC TIRE 235

COS </>

FIGURE 4.30 Laws of change of R-cos $ at a point of the lattice.


(a) During panlographing.
(b) During cord stretch.
(c) During ply slippage.

4.7. Analysis of Manufacturing Methods by R vs cos <j>


Charts
The fact that the R vs. cos <j> presentation of changes of cord path en
ables the character of the changes to be understood, and the fact that the
R vs. cos <f> representation of the actual cord path together with the R vs. X
diagram for the surface of revolution on which the cord path lies com
pletely defines the cord path in space, enables a system of cord path di
mensions, and changes under manufacturing conditions to be set out.
Consider three sets of data expressed in the form of five graphs or tabu
lations, as shown in figure 4.31. The graphical form is easiest to under
stand, although the numerical form enables the required precision to be
met.
The curves of figure 4.3 1 labeled I show the cord path characteristics or
R vs. cos </> information, curve a, on the tire building former, which is a
surface of revolution defined by the R vs. X statement determined by the
generator of the surface, curve b.
The curves labeled II show the changes due to trellissing or pant-
ographing, changes in cord extension, and any slippage of the plies at the
trellissing pivot points, during manufacture, curve c.
The curves labeled III show the final cord path characteristic, or R vs.
cos <>, curve d, in the finished tire casing of shape defined by its R vs. X
information, shown in curve e.
Given any two sets of information out of the three, I, II, III, the other
can be found explicitly, or, given any four of the five items of information
(a), (b), (c), (d), (e), the fifth can be found exactly.
It is possible that a bad choice of the given data will result in the re
maining calculations taking on an impractical form. This is not a limita
tion or fault of the method, but rather the reverse—it shows the limitations
of what can be done in the process of manufacture.
The calculations do not involve difficult mathematical functions, in-
236 MECHANICS OF PNEUMATIC TIRES

Cos.fr

FIGURE 4.31 Representation of R-cos <t> method.


I Original plied materiml on former.
II Shaping Process.
HI Final lire cord path.

complete integrals or mathematical singularities, as would a method in


which the length of a path is determined analytically from the coordinates.
This method is, as mentioned earlier, one of step by step resolution of the
cord path into the coordinates of the surface, analogous to navigation.
The details of the application of the method outlined here, including the
choice of the incremental steps along the path of the cord so as to avoid
systematic cumulative errors, along with examples of applications, are
given in reference [1]. This reference also gives information on the R vs.
cos <j> characteristics of paths other than the pantographing and geodetic
cases.
In addition to the above uses, the R vs. cos <J> method of presenting in
formation can be used as a diagnostic tool to investigate why a given tire
does not conform to the expected cord path. The departures from the ex
pected path can be interpreted in the light of the observed departures of
the R vs. cos <j> from the normal characteristic. Cord crowding due to in
sufficient spacing in the initial material, cord extension, tight cords, and
cord or ply slippage each have a specific effect on the chart.
It is a general truth that the most likely change during the shaping proc
ess is pantographing, since departures from pantographing and uniform
stretch are caused by local effects, such as the tire contacting the mold at
one region before another and the subsequent movement of parts of the
casing being restricted by this contact with the mold.
The fact that the high modulus cords are embedded in a low modulus
matrix which adheres to them is the reason why pantographing, corre
sponding to cos <(>/R = C1, occurs most readily. The low modulus matrix
provides a low resistance to the shearing movements involved in pant
ographing, while the adhesion between the high modulus filaments or
cords and the low modulus matrix or rubber resists slippage of one ply
across another, and ensures pin-jointing or pivoting at filament inter
section points between two adjacent layers of opposite bias.
Since a cylindrical former belongs to the group of axisymmetric surfaces
STRUCTURE OF THE PNEUMATIC TIRE 237

which are developable, so that they can be made from a flat surface with
out internal distortions in the plane, then plies laid together on the flat,
wrap round the cylinder without pantographing. If the form is of any
shape other than a cylinder, including those regions of a so-called flat (cy
lindrical) former which depart from a true geometrical cylinder, pant
ographing will occur appropriately.

4.8. Mechanism of Load Carrying—Infinitely Flexible


Membrane
To understand the fundamentals of the mechanism by which a pressure
inflated structure carries a load, consider cases of infinitely thin and flex
ible membranes. The flexibility is to be such that the forces set up by
bending of the film are so low that they can be ignored. Thin soap films
are often quoted as examples of such membranes.
Consider a tube of these characteristics, held in friction-free contact be
tween two parallel flat plates whose distance apart is such as to compress
the tube as in figure 4.32. At any point in the wall of the tube not in con
tact with the plates, an element of the surface is indistinguishable from the
surface of a complete tube of the same material and characteristics free of
contact with any plates, but of diameter 2r, the separation distance of the
two plates in the first case. It follows that the tension in the free wall of the
tube held compressed between the two plates is t = pr, from elementary
equilibrium. Since this holds right up to the actual point of contact with a
plate, it follows that the free wall is truly semi-circular of constant radius
of curvature r, and it meets the plates tangen tially at the four contacting
points. The total reaction on plane AB is p · (AB), and the tension in the
free walls at A and B is each / = pr, and also AB = 2r + 2a where 2a is the
length CD of the film on the plate. Now, from equilibrium arguments, the
total reaction on CD = total reaction on AB minus the resultant, in the di
rection normal to CD, of the pressures on the curved walls A C and DB.
This leads to
p · (CD) = 2p(a + r)- 2pr = 2pa.
This is exactly what is to be expected when it is noted that the gas pressure
acts in a direction normal to the plate CD, and to the flexible membrane in
contact with the plate. Furthermore, as this surface of the tube is parallel
to the plate CD, the tension in that part of the tube does not matter since it
has no component normal to the plate. In the case just discussed, the ten
sion in the wall just as it meets the plate tangen tially is t = pr, and this

FIGURE 4.32 Flexible cylindrical tube compressed between two parallel flat plates.
238 MECHANICS OF PNEUMATIC TIRES

must be resisted by the tension in the film across the plate which is in fric
tion-free contact with the plate, so that the tension in the film or mem
brane in contact with the loading plate ist = pr. If the friction between the
loading plate and tube were not zero, then the tension in the membrane
across the plate would be modified according to the friction laws pertinent
to the contacting surfaces, and possibly the manner of approach of the sur
faces, as well as the elastic moduli of the tube wall and platen. But in any
event this would not alter the load carried normal to the plate.
The main point to be noted is that as compared with the free tube, the
use of the tube to support load on the parallel plates reduces the tension in
the free walls of the tube, and in the case discussed the load carried is pre
cisely equal to the gas pressure times the contact area between tube and
plate.
If the wall of the tube could not be assumed infinitely flexible, then the
bending stiffness and/or wall shear rigidity would cause the load to be
supported by the platen to be greater than the pressure times the actual
contact area. The stresses in the wall would be modified near the contact
region, locally the wall curvature might vary from the semicircular arc,
and the tension in the membrane across the contact might be different. At
this moment these effects are not relevant to the matter directly under dis
cussion but are pertinent later.
It will be noted that in the above case, if the material is inextensible as
well as infinitely flexible, the overall periphery of the tube remains con
stant so that 2irr + 4a = 2-nR, As the load carried is 2ap = L, it can be de
duced that the plates approach towards each other by an amount

up

or put another way, at a deflection x the load carried is

LT = 2ap
1

where L is the load per unit length of the tube.


Now consider the case of a complete spherical membrane or balloon,
flexible yet inextensible, compressed between two parallel flat plates. The
principal radii of curvature of the surface are r1 = radius of curvature of
the cross section in a plane radial to the axis of symmetry of the surface,
and r2, the distance from the surface to the point where the radius r1 ex
tended meets the axis of symmetry of the surface, as in figure 4.33.
As the surface of that part of the balloon or membrane in contact with a
plate is flat, r2 at a point immediately adjacent to the edge of the contact
patch is infinite and we have from the equation of normal equilibrium (see
fig. 4.34).

Since p is constant throughout the interior not in contact with the wall, in
cluding the region of the equator between the two plates where the mem
STRUCTURE OF THE PNEUMATIC TIRE 239

FIGURE 4.33 Sketch of spherical element and nomenclature.

brane is nearly spherical and has a tension t, approximately the same in


both t1 and /., directions, then
t, 2t 'i " '' ^_

where R is the radius of curvature of the spherical wall. Now t1 = t be


cause, at any point on the curve representing the cross section in the radial
plane (fig. 4.34) nothing contacts the membrane except gas inflation or the
plate, aoAsejust at the point where the membrane leaves the plate, t1 = t
and so^T= R/5>as in figure 4.34.
Therefore tne radius of curvature of the cross section at the point in the
membrane where it touches the plate is R/2 and the plate is tangential to
the membrane surface arc at this point. The surface follows a transition
curve from this point to the main region of the sphere, as in figure 4.34.
Because of these facts the radius b of the actual contact is less than a, the
radius of the intersection of the sphere and plane in figure 4.34.
As stated earlier, the load on the plate is equal to the product of the in
ternal gas pressure and the actual contact area of the membrane on the
plate. The only other loading which could come on to the plate is a shear

FIGURE 4.34 Membrane in contact with plate.


240 MECHANICS OF PNEUMATIC TIRES

(a) (b)

FIGURE 4.35 Toroidal membrane on cylindrical rim.

loading within the thickness of the membrane around the perimeter of the
contact patch, and this could only arise as a consequence of bending
Stresses in the membrane. Because of the initial assumption of infinite
flexibility, these stresses cannot be present in the problem as discussed.
It also follows that the two contact patches are equal in area and that
their midpoints are on the same normal to the flat plates.
Now consider a toroidal or tire-like structure of the same material as the
previous balloon membrane but with a rigid tubular rim for the central
zone or bore of the toroid, as in figure 4.35a. Assume that the junction be
tween the thin flexible membrane and the rigid tubular rim or base has
zero bending rigidity.
Inflation of the structure puts tensions in the membrane and it takes a
shape as determined by equilibrium and compatability conditions. The
membrane tensions are resisted by reactions at the edge of the tubular rim,

FIGURE 4.36 Perspective sketch of toroidal shell contact.


MECHANICS OF PNEUMATIC TIRES 241

and for our present purpose these can be discussed in terms of two com
ponents viz, radially outward tension and tension in an axial direction
(that is parallel! to the axis of rotational symmetry or rotation of surface
generators) at each point around the edge of the rim.
If a flat plate is pressed against the membrane while the structure is sup
ported by the rim a reaction will develop between membrane and plate in
the manner already discussed for the sphere and the cylinder, as in figure
4.35b. By the same arguments the load will be equal to the product of the
actual contact area and the inflation pressure. The contact area will be ap
proximately but not exactly elliptical, but its geometrical shape is of im
portance only foflHe discussion of tfie point about to be made.
The question now arises, how does the reaction develop at the rim? The
air pressure is uniform and the rim base width is constant, and it follows
that the resultant of the air pressure reactions on the rim is zero even when
the plate is pressed against the membrane.
Consideration of the structure shows that the only possible way in
which the reaction can develop at the rim is by the cKanges in magnitude:
and direction of the membrane stresses at their points of attachment to the
run, in the region of the membrane near the point where the plate is
pressed against it.
As can be seen from figure 4.36 the curvature of the wall of the mem
brane increases in the region between the loading plate and the adjacent
rim. Hence, because of the increasgd curvature the membrane stresses in
this region are lower than elsewhere in the membrane walls, since t = p · r.
This deflection also causes the membrane to distort locally, increasing the
angle between the direction of the wall and a line normal to the plate from
the rim; this is true whatever the cross-sectional shape, shown in figure
4.37. This increase nf angle reduces algebraically the cosine of the angle
between the wall and the line of action of the applied load on the plate.
The net effect of the reduced tension and reduced component at the de
flected region is to develop the required reaction. The rim, in effect, hangs
in the tensions of the undeflected walls as shown in figure 4.38. i he radi-
ally outward components of the wall tensions are greater in the undef
lected regions than in the deflected region. The system of load transmis
sion is analogous to that of a cycle wheel where the hub hangs by the steel
wire spokes from the top of the rim, which is loaded at the bottom as illus
trated in figure 4.39.
The above analysis assumes that the membrane structure is so flexible
that the load which would be carried by it when uninflated is trivial.
If, as is usually the case, the structure had some bending stiffness, this
would contribute some load support, and to do this it would need to de
velop bending stresses in the membrane, particularly around the edge of

FIGURE 4.37 Cross sections offigure 4.37 showing deflections ofsidewalls which reduce the
tension component radially outward at the inner cylinder edge.
242 MECHANICS OF PNEUMATIC TIRES

II Mt
FIGURE 4.38 Polar plot of radially outward component of wall tension of membrane loroid
on inner cylinder.

the contact patch where it contacts the loading platen. Bending stresses
"may also, but not necessarily, develop at the rim edge, where the presence
of a flange on the rim would, of course, set up bending moments depend
ing on its shape and the bending stiffness of the membrane at that region.
The objective of this discussion of some apparently academic cases is to

Spoke Tension

Spoke Tension

Rim


Ext. Lood

FIGURE 4.39 Load on hub of cycle wheel is carried by change of tension in spokes.
STRUCTURE OF THE PNEUMATIC TIRE 243

FIGURE 4.40 Effect of casing bending stiffness on contact pressure.

get principles clear, because the actual tire case involves them and some
further complications, particularly in the bead region because of the steel
bead coils.

4.9. Mechanism of Load Carrying—Tire Structure


In an undeflected tire the cords are tensioned by the excess pressure of
the inflating gas over the external or atmospheric pressure. The tire casing
takes up its equilibrium shape, which is determined primarily by the cord
paths, perhaps modified somewhat in local regions by the presence of ex
tra rubber, ply turn-ups, flippers, filler, etc. For clarity and ease of dis
cussion we will ignore these features and their effects and start from the
simple theoretical equilibrium shape.
As the tire is pressed against a flat roadway the tread rubber is com
pressed and at the same time the tire casing locally loses its axial symme
try and takes on a substantially flattened contact patch. If there were no
tread rubber on the tire the casing would be flat over the area actually
contacting the ground.
Let us continue our examination of the problem assuming that, for the
present, there is no tread rubber. The cords in the area of the casing in ac
tual contact with the ground will also lie in a flat plane parallel to the
ground plane and hence, viewed along any line parallel to the ground
plane, the line of the cord is straight. It follows that the tension in the cord
bears no relation to the internal inflating gas pressure, in this particular re
gion of te lire. To make this point emphatically clear it will be noted that
the"basic law determining the relation between tension in the casing cords
and the internal gas pressure which it resists is based on the simple laws of
statical force resolution, and in such cases a path of infinite radius of cur
vature results in zero resultant force opposing the gas pressure. The ten
sion in the cords across the flat part on the contact patch is therefore deter
mined primarily by the cord tension transmitted from the adjacent free
wall ofthe tire, modified by the effects of the transition curvature around
the perimeter of the flat contact patch. It also follows that the contact pres
sure between the tire casing and the ground will be equal to the inflation
pressure, modified around the edges of the contact patch by the extra pres
sures set up by the bending stresses within the transition zone, as shown in
figure 4.40.
In the case of practical tire designs the presence of tread rubber of dif
fering thickness, of tread pattern design, and such factors as the bending
stiffness caused by the multiple layers of cord cause the actual tire contact
244 MECHANICS OF PNEUMATIC TIRES

I 2
INS DEFLECTION

FIGURE 4.41 11.00 X 20 tire deflection data over a range ofpressures.

pressure to be locally greater than the inflation pressure, and in fact to dif
fer in different parts of the contact patch. The contact pressure at the sides
of the contact patch, under the shoulders of the tire tread, is often higher
than the general contact pressure because of the reaction necessary to de
velop the bending stresses in the transition zone around the contact area.
This is most marked if the shoulders of the tire tread are thicker than the
tread at the crown of the tire.
For these reasons the actual load carried by the tire is higher than the
product of the overall contact area and inflation pressure. The extra reac
tion is carried by the tire acting as a structure.
The load carnedby thesetwo processes seems to be linearly additive.
The load-deflection data of figure 4.41 can be analyzed by redrawing the
various load-deflection curves for the range of inflation pressures as a lat
tice plot, as in figure 4.42, In this plot the load-deflection curve for a given
pressure is started from a zero determined by the value of the inflation
pressure, the zeros being located along the abscissa on a linear pressure
scale.
Points on each curve for a given combination of load and deflection are
taken from figure 4.41 and are joined to produce a line of constant deflec
tion. This process produces a family of constant deflection curves laid,
across the original family of constant pressure curves. For any combina
tion of pressure and deflection, the load developed can be read off. Inter
polation at intermediate values is easy.
The constant deflection lines can be extrapolated to produce a curve for
zero inflation pressure, shown as a dashed line in figure 4.42. The load at
any specified pressure and deflection is seen to be the linear sum of the
part dependent on the inflation pressure and the part due to the structural
stiffness, or the stiffness at zero inflation pressure.lt should be noted that
the slope of a constant deflection line, shown chain dotted, is an effect ive
area and is of the order of the contact area, as would be expected from the
discussion above.
The stiffness of the structure can be expressed as an equivalent effective
pressure, the value of which is determined by extrapolation of a constant
STRUCTURE OF THE PNEUMATIC TIRE 243
— Constant Pressure
lOOpsi — do Extrapolated
-Constant Deflection
80 do Extrapolated
60

INCHES

FIGURE 4.42 Lattice plot of 11.00 x 20 tire data offigure 4.42 showing both load-deflection
curves at various pressures and also load-pressure curves at various deflections, thus separat
ing pneumatic load and direct structural load.

deflection line through the line for zero inflation pressure. The horizontal
intercept from the foot of the ordinate of the point where the given deflec
tion line cuts the zero inflation line, to where the given deflection line cuts
the zero load base line, determines an effective inflation pressure equiva
lent to the stiffness of the structure as shown in figure 4.43. Construction of
curves of this type is described in references [2-4].
So far, the development of the load at and in the immediate vicinity of
the tire-ground contact has been discussed. The load must be transmitted
to the wheel and this involves not only the transmission of forces between
bead and rim but also the transmission through the structure. The area of
the wheel rim base acted upon by the air pressure is axially symmetrical
and is a cylinder of constant length corresponding to the rim width, so the
uniform inflation pressure cannot produce a resultant force on the rim.
To postulate how the force set up by the inflation pressure acting
through the flattened contact region can produce stresses in the structure
having a resultant equal to the load carried, consider two mechanisms of
force transmission acting in parallel.
First of all, consider a band approximately the width of that part of the
tire tread and casing in contact with the ground, and with the same prop-

too psi

80

40

20

INCHES

FIGURE 4.43. Line of constant deflection extrapolated to estimate structural stiffness as an ef


fective inflation pressure.
246 MECHANICS OF PNEUMATIC TIRES

erties and construction as that region, extending around the whole periph
ery of the tire. This is shown in figure 4.44. At all points the inflation pres
sure acts on this band, and, at all places except that pan in contact with
the ground, it is curved and so the tensions set up by inflation pressure re
sist and equilibriate the inflation pressure. In that portion in contact with
the ground the inflation pressure forces are transmitted by compression
through the band without producing a resultant reaction on the band or
ring. The net result is that the absence of forces on this pan of the band
leaves the forces due to inflation pressure on the top sector of the band un-
resisted by an opposite force on the bottom sector, while at other parts of
the tire outward forces have zero resultant, as shown in figure 4.44a.
The resultant upward force on the upper half of the ring, being greater
than the resultant downward one on the lower half, causes the side wall
tensions in the upper half to be greater than the lower half, as shown in
figure 4.44b, and this force pulls the bead coil against the base of the
wheel rim above the contact area, thus transmitting the upward force to
the wheel, as shown in figure 4.44c.
The other mechanism of force transmission from the ground to the
wheel is analogous to that discussed earlier for the cases of a cylinder and
a sphere, where the deflection of the walls at the contact region lowers ten
sion forces in the walls in that region, and here, because the bead coil has
a high modulus, it bears on the base of the rim via the ply material around
it with a force just sufficient to make up the total load, as shown in figure
4.44.
Both mechanisms lead to the same kind of transmission of force from
tire to rim—the wheel rim hangs in the bead coils, which in turn hang in
the tensioned casing cords which have lower tension in the contact region
than elsewhere.

FIGURE 4.44 Load transfer from ground to wheel by effects of inflation.


STRUCTURE OF THE PNEUMATIC TIRE 247

f In addition to the load carried by a combination of these two mecha


nisms, some load is carried by the structure of the tire in the region of the
contact, as would be expected if it were a solid tire. This load is carried by
the structure of cords and rubber from the contact patch to the rim as a
compression in the wall held stable by the inflation pressure and curva
ture, but its contribution is small. The structural load may also be trans
mitted via the bead ring system already discussed as a mechanism for the
inflation load.
The true stress pattern in the tire structure is the sum of these several
effects. An analysis of the problem requires the application of numerical
methods; the finite element technique is the most convenient method of
obtaining an estimate of the total structural stiffness. Two points will be
made, however. The true estimate of structural stiffness requires the devel
opment of true finite displacement analysis—the analysis of lattices with
substantial deformations so that the compatibility of the framework is not
violated, even where the framework is distorted substantially from its orig
inal state of axial symmetry. The second point is that even when satisfac
tory structural stiffness and load carrying capacity analyses have been es
tablished, the use of finite element models for stress estimation must be
used with care so that the true peak stresses in the structure are estimated
accurately; these are the ones which determine failure starting points and
fatigue life. The complexity of the problem is illustrated in figure 4.45.
The stress is difficult to obtain at points such as A, the free end of a cord in
a region where the stress in the low modulus matrix is changing rapidly,
and B, the closest points on the surfaces of nearby crossing cords. Element
sizes need to be small to cope with this.

FIGURE 4.45 Points of high stress in breaker and other regions.


248 MECHANICS OF PNEUMATIC TIRES

References
[1] Gough, V. E., Cord path in tyres. Trans. I. R. I. 30 (1), T20 (1964).
[2] Gough, V. E., Tyre to ground contact stresses, Wear 2, 107-126 (Nov. 1958).
[3] Cooper, D. H., Radial stiffness of the pneumatic tyre, Trans. I. R. I. 40 (1), T58 (1964).
[4] Gough, V. E., Tyres and air suspensions, Advances in Automobile Engineering, Tidbury,
ed., p. 59 (Pergamon Press, 1963).
Chapter 5
CONTACT BETWEEN THE TIRE
AND ROADWAY

Alan Browne,1 K. C. Ludema,2 S. K. Clark3

5. 1 Introduction 250
5.2 Contact Area 25 1
5.2.1 Experimental Measurements 251
5.2.2 Theory 261
5.3 Normal Pressure Distributions 271
5.3.1 Experimental Measurements 271
5.3.2 Theory 282
5.4 Tangential Pressure Distributions 292
5.4.1 Experimental Measurements 292
5.4.2 Theory 299
5.5 Slip Between Tire and Roadway 303
5.6 Tractive Forces on Dry Pavements 308
5.7 Tractive Forces on Wet Pavements 320
5.8 Tractive Forces on Snow and Ice 349
5.9 Tractive Forces on Unpaved Surfaces 356

' General Mown Research Laboratories, Warren. Michigan


Department of Mechanical Engineering, University of Michigan. Ann Arbor, Michigan
Department of Mechanical Engineering, University of Michigan, Ann Arbor, Michigan

249
250 MECHANICS OF PNEUMATIC TIRES

5.1 Introduction
The purpose of this chapter is to examine both the experimental data
and the appropriate mathematical models for three characteristics of a
rolling elastic tire. These characteristics are:
a. The area of contact between the tire and road.
b. The slip or relative motion between the tire and road.
c. Normal and tangential contact stresses between tire and road.
Since the forces needed for vehicle support, guidance and maneuvers all
arise in the tire contact area study of these characteristics and application /
of the resulting theory should improve both the mechanical properties of
tires and the control of vehicles. These characteristics are all influenced by
the operating parameters of a tire, such as its inflation pressure, its rolling
velocity, its path of rolling relative to its midplane, its load as well as its
size and shape/For this reason it is quite difficult to give a complete defin
itive description of all possible interacting factors, and for the most part it
will be necessary to rely on rather insufficient experimental evidence
which merely indicates trends and magnitudes, as well as some relatively
simple theoretical ideas which from time to time may prove helpful in ex
plaining the general characteristics of observed phenomena.
In addition to the influences which are listed above, it is obviously true
that the characteristics of contact between a tire and roadway depend as
well on the particular type of roadway surface to be studied, as well as on
any possible contaminants between the tire and roadway. For the most
part one is interested in the influence of water in relatively small depths,
and its effect can be marked on contact processes, particularly at high ve
locities. Roadway surfaces usually are constructed to be relatively smooth
compared to tire tread patterns and there is little if any information avail
able in the literature on the influence of road roughness on contact charac
teristics.
The quantities of interest here are shown in the sketch of figure S.I.

FIGURE 5.1. Total reaction from ground to lire expressed as the combined effect of two
forces, one normal and one tangential to the ground plane.
CONTACT BETWEEN THE TIRE AND ROADWAY 251

AIRCRAFT i rtC
32»88
8 PLY -RATING 86 ps.
6000 fe \*rt
1-3/4" Vert. Defl

FIGURE 5.2. Contact patch of a typical aircraft tire.

5.2. Contact Area


5.2.1. Experimental Measurements
Influence of Tire Variables
The shape of the contact area between the tire and roadway depends on
the tire cross section shape and structure. For example, the contact area
between an aircraft tire and a Sat surface usually appears to be nearly el
liptical in shape. This is illustrated in figure 5.2. This is generally the kind
of shape associated with a tire which has little if any external tread or
shoulder region, but is primarily a toroidal carcass with small additional
tread rubber.
For an automotive tire a somewhat different set of relationships exists
due to the fact that the usual construction involves the use of a relatively
heavy tread, particularly in the shoulder region. In this case any signifi
cant contact spreads over the entire width of the tire between shoulders so
that the contact area tends to have essentially straight parallel sides, and
the width of this contact area is nominally independent of tire deflection.
A typical contact area is shown in figure 5.3.
Data relating gross contact area to tire deflection is shown in figure 5.4
for a typical set of automotive tires of different size and manufacture. It is
seen from this that the relationship between tire deflection and gross con-
252 MECHANICS OF PNEUMATIC TIRES

8:00x14 Automotive Tire


Four ply Royon Bias Ply Construction
24 psi inflation 1.25" Deflection
1350 Ibs load

FIGURE 5.3. Typical contact area for an automotive lire.

tact area is nearly linear, accounting for some experimental scatter in the
data. Similar results have been reported by Hadekel [1]2 previously.
I Experimental evidence indicates that tire deflection is the most impor
tant variable governing the area of contact between the tire and roadway.
If inflation pressure and load are simultaneously varied so as to maintain
Constant tire deflection, the contact area of the tire will remain effectively
constant. 'This conclusion was previously reached by Michael [2] on the

Figures in brackets indicate literature references st the end of this chapter.

9
50
0
fi
740 4i
2 35 C
(X 0 .
< 30 I, 1 A

0 g
<1
p • 8: 00x14 BIAS F 1.Y
o 7: 50x14 BIAS F 'LY
a A. 5: 90x15
n BIAS F >LY
§ io o 7: 50x14 STEEL RADI HL
£ 5 • 1 55x15 TEXTI .E RA 3IAL

.25 .50 .75 1.00 1.25 1.50 1.75 2.00


TIRE DEFLECTION - INCHES

FIGURE 5.4. Gross contact area vs. tire deflection.


CONTACT BETWEEN THE TIRE AND ROADWAY 253

basis of aircraft tire data, but is also shown here in figure 5.5 from unpub
lished data [3]. This tends to lend credibility to the inextensible membrane ,lr
or shell bending models for contact phenomena.
There is not too much evidence available in the literature concerning
the role of curved surfaces in forming contact areas with pneumatic tires.
For example, the contact of a pneumatic tire with a cylindrical surface is
of considerable interest due to the large amount of indoor tire testing car
ried out on cylindrical steel roadwheels. Some preliminary data is avail
able on this effect from the work of Dodge [3] and this is shown in figure
5.6, where the variations of contact patch length with surface radius are
shown for two different automotive tires.
The contact area of a slowly rolling tire may be slightly different from
that of a standing tire, but it is not clear that such differences would be
large. In some cases carcass deformation would be slightly different in
standing as opposed to pure rolling due to the presence of friction between
the roadway and the wheel. Experimental data on this point has been ob
tained by the U.S. Army Engineer Waterways Experiment Station [4] for a
single truck tire. This data is summarized in table 5.1.
There is only limited data available on the influence of velocity on the
area and shape of the contact region between a tire and roadway. Seitz [5]
gives footprint data taken on a roadwheel. In figure 5.7 is shown the con
tact area for a textile belted tire at various speeds at constant load, while
similar data for a steel belted tire and a bias-ply tire is shown in figures 5.8
and 5.9 respectively. Figure 5.10 shows contact areas as a function of tread
radius, while figure 5.11 shows the bias-ply tire contact area again as a
function of velocity. Figure 5.12 gives a similar presentation for a belted

10

in a i i
Ul O f|-, — B
0 ci o r
• o o
o ' " o' " ' 4 i
, ° <>

7
4> 7:50x14 STEEL RADIAL
«9 CONSTRUCTION
z
w 1 7:50x14 BIAS PLY
• CONSTRUCTION

Q.
ALL OAT;V TAKEN AT t" 0 EFLECTION
9

1 1 1 1 I
10 20 30 40 50
PRESSURE (PSD

FIGURE 5.5. Plot of contact patch length vs. pressure for two automotive tires affixed
deflection.
254 MECHANICS OF PNEUMATIC TIRES

I r- -r- —r- —i r
B ALL DATA AT 28 PSI INFLATION
10
1" DEFLECTION
• BIAS CONSTRUCTION .
o BELTED CONSTRUCTION

1.5" DEFLECTION
• BIAS CONSTRUCTION —
D BELTED CONSTRUCTION

, ,
7248
I
24
CURVATURE (inches)"1
i
12
i
6
RADIUS OF CURVATURE (inches)

FIGURE 5.6. Plot of contact patch length vs. roadwheel curvature for two different tires and
two different deflections.

tire whose contact area is a function of velocity. These measurements were


made on a cylindrical roadwheel, but still give a valuable indication of the
general influence of velocity on contact area. While the trends are not
completely clear, it appears that increasing velocity of travel generally
tends to increase the contact area slightly, all data being taken at constant
load. It is known that increasing velocity may cause the center of normal
pressure to move slightly forward in the contact area, since the power loss
characteristics of the rolling tire may result from such a forward shift. This
apparently is not accompanied by a geometric shift of the actual contact
area, according to the data of Seitz.
Finally, some photographic information is available on the nature of an
actual contact patch under rather slow cornering conditions. This was ob
tained by photographing through a glass plate as the tire is run over the
plate. A still picture from such a movie is given in figure 5.20 showing typ
ical cornering distortion of the contact patch.
Influence of the Environment
I Operating environment can have a significant effect on both area and
shape of the contact region.
CONTACT BETWEEN THE TIRE AND ROADWAY 255

TABLE 5. 1 Static and rolling contact patch areas and dimensions (slow rolling conditions)
Inflation Load Deflection Contact area
Test
condition psi pounds inches % length width area
Stationary 60 4500 1.56 17.3 11.4 6.4 67
Rolling 60 4500 1.56 17.3 11.4 6.6 67
Stationary 60 3000 1.12 12.5 9.9 6.0 52
Rolling 60 3000 1.12 12.5 10.2 6.0 58
Stationary 60 1500 0.50 5.5 7.7 4.6 29
Rolling 60 1500 .50 5.5 7.8 4.8 31
Stationary 30 4500 2.22 24.6 16.0 8.2 125
Rolling 30 4500 2.22 24.6 15.8 8.5 115
Stationary 30 3000 1.75 19.4 12.4 7.4 83
Rolling 30 3000 1.75 19.4 12.4 7.1 81
Stationary 30 1500 0.88 9.7 9.9 5.6 47
Rolling 30 1500 .88 9.7 9.2 5.3 43
Stationary 15 4500 5.00 55.4 20.0 9.3 188
Rolling 15 4500 5.00 55.4 20.2 10.5 161
Stationary 15 3000 2.75 30.4 16.1 8.2 130
Rolling 15 3000 2.75 30.4 15.5 8.2 115
Stationary 15 1500 1.31 13.1 10.5 6.6 63
Rolling 15 1500 1.31 13.1 11.0 6.2 63

When fluid contaminants such as water, oil, slush, or mud are present
on the pavement surface, contact area between tire and pavement is re
duced due to the persistence of a fluid film in portions of the formerly dry
contact area. This resistance of the fluid to expulsion from the interface is
tied to the viscous and inertia! forces developed in the fluid as it is wedged
and/or squeezed between tire and pavement surfaces. The reduction in
traction that occurs on a fluid contaminated pavement is indirect evidence
of this loss in dry contact area. /
Direct evidence of this effect has been obtained [6, 7] by using a glass

mm —1—
80 1 <•*• =T3
| .V = H9.8(km/h)
1 1 V = 60.5(km/h)
70 , -E&N
-V = 30.0(km/h)
60 1

SO f_J ji
I|S
40
30 '!\'i .
1 pL=l.7 (kp/cm2)
(inflation)
20 | T P=360(kp)
i
10
q 33
}0 TO
^*'*>

LEACHNG EDGE
50 30 10 10 30 50
.-^
70mm
J^

TRAILING EDGE
(load)

FIGURE 5.7. Contact area for a textile belted tire at various speeds, taken at constant load.
256 MECHANICS OF PNEUMATIC TIRES

V«30(km/h)
V»60(km/h)
V* 120 (km/h)

p.*l.7(kp/cm2)
(inflation)
P«360(kp>
(load)
-90 -70 -50 -30 -10 ] 10 30 50 TOTRAILING
90mm
LEADING EDGE EDGE

FIGURE 5.8. Contact area for a steel belted tire at various speeds, at constant load.

mm
80 V = 120.2 (km/h)
6l.3(km/h)
70 V = 30.2 (km/h)
60
50J
40 Pi'l.7(kp/cmr
(inflation)
30
P=360(kp)
20 (load)
10
0
-70 -50 -30 -10 10 30 50 70mm
LEADING EDGE TRAILING EDGE

FIGURE 5.9. Contact areafor a bias ply lire, with 37° crown cord angle, at various speeds, at
constant load.

P*37D(Kp)dOOd)

-80 -60 -40 -20 20 40 60 80


LEADING EDGE TMUUNGEME CROWN PROFILE

FIGURE 5.10. Contact areas for a bias ply tire with 37° crown cord angle as a function of
tread radius, taken statically at equal loads.
CONTACT BETWEEN THE TIRE AND ROADWAY 257

I20.3(km/hr)
V-60.7(km/hr)
V30.5(km/hr)

1.7 (dp/cm2)
(inflation)
P'360(kp)
-70 -50 -30 10 30 50 70mm (load)
LEADING EDGE TWUUNGEDSE

FIGURE S.I I. Contact area of a normal tire as a function of velocity, at constant load.

plate embedded in the roadway surface. Photographs such as Fig. 5.13


taken through such a facility clearly show the persistance of a fluid film
(grey region) in portions of the formerly dry contact area. The arrow in
dicates direction of tire travel. Just how much of the dry contact area
(dark area) will remain depends on tire, pavement, fluid, and vehicle fac
tors in the manners shown in Table 5.2. As an example, Fig. 5.14 displays
the effects of speed and groove depth on the area of dry contact for a bel
ted bias tire as determined from glass plate photographs. The shape of the
remaining dry contact area also depends on many factors. This is a more
complete description than the simplified view implicit in the three or four
zone concept of contact area under wet conditions as discussed in [8] and
[9].
One dimensional theoretical models based on the three or four zone
concept assume a steady decrease in fluid film thickness from the front to
the rear of the footprint region. The glass plate photograph in Fig. 5.15
shows that this is not always valid. Here a fluid film (grey region) occupies
the front, central and rear portions of the footprint, with traction-produc
ing contact (dark regions) occurring only under the side ribs. The arrow
indicates the direction of tire travel. Whether the fluid film will persist
along the centerline as illustrated in Fig. 5.15, as opposed to the manner
predicted by the consecutive zone concept (Figs. 5.13 and 5.16), is deter-

BJU
90
80

V>— -V'6O(km/h)
TMILMGEDeE 70 * =- "».
V«l20(km/h)
60 [
\
50
40
30
i
1 I
A
Vpi-2.5(kp/cmt)
i
20 (inflation)
10 2
0
^—
-20
^ I
0
-"L aP*6O(kp)
20n m(lood)

FIGURE 5.12. Contact area of a steel belted tire at high inflation pressure and extremely light
load, at constant vertical load.
258 MECHANICS OF PNEUMATIC TIRES

FIGURE 5.13. Footprint of a steel belted radial tire;free rolling at 45 m/sin 2.0mm of water
m-
mined by whether the fluid inertial forces are large enough to bend the tire
surface inward along the footprint centerline. Tires with rigid carcass con
struction, high inflation pressures, and large tread element size will tend to
conform to the consecutive zone concept at low to moderate speeds on rel
atively smooth pavements having little microtexture and rounded mac-

Woter Depth : 0.080 in


Tire: Bios Belt
Inflation: 28psi
Load: 1340 Ib
Free Rolling
6 Grooves , 0.375 in wide

30 40 50 60 70 80 90
VEHICLE VELOCITY, mph

FIGURE 5.14. Effect of groove depth and velocity on ground contact area [17].
CONTACT BETWEEN THE TIRE AND ROADWAY 259

TABLE 5.2 Factors affecting percentage ofdry contact area

Change in
fractional dry
contact area with
increase in
Variable Factor factor
Tire Inflation pressure Increase
Carcass rigidity Increase
Width of grooves in tread pattern Increase
(number held constant)
Number of grooves in tread pattern Increase
(width held constant)
Groove depth (width held constant) Increase
Size of distinct tread elements Decrease
Pavement
Microtexture • Increase
Macrotexture Increase
Porosity Increase
Fluid
Viscosity Decrease
Density Decrease
Depth Decrease
Vehicle
Speed Decrease

DIRECTION
OF
TRAVEL

1
FIGURE 5.15. Footprint of a bias ply tire;free rolling at 45m/s in 2.0mm of water [7J.
260 MECHANICS OF PNEUMATIC TIRES

rotexture and covered by thin fluid films. Tires with flexible carcass con
structions, low inflation pressures, and narrow, widely spaced grooves
operating at high speeds on pavements with little macrotexture covered by
thick fluid films will lose dry contact principally along their footprint cen-
terline. Figure 5.17 contains three photographs showing how in this latter
case, the fluid film region will first grow along a narrow band through the
center of the footprint and then extend laterally toward the side edges of
the footprint as speed increases. Dark bands on the photographs are Moire
fringes produced by the optical technique described in [10]. They repre
sent equal increments in fluid film thickness and, thus, contour lines of the
deformed tire surface. White tread stock was used to create good optical
contrast. The arrows again indicate direction of tire travel.
Even when fluid contaminants are present in insufficient quantities to
flood the major groove network in the footprint region and bend the tire
surface upwards from the pavement, they can still reduce contact area and
traction by persisting under individual tread elements. In experiments per
formed by Roberts [11] on the approach of a rubber sphere onto a lubri
cated flat glass surface, Fig. 5.18, and more recently by Field and Nau [12]
with a cube of rubber, a pocket of fluid was shown to be trapped between
the two surfaces. "Dry" contact first occurs only around the perimeter of
the surface of the rubber element. In the absence of any texture from both
surfaces, such as would exist with a tread element without sipes on a pave
ment lacking microtexture, the trapped pocket of fluid persists for a con
siderable length of time [12]. This event has recently been analytically
modeled [13] and studied in terms of its relationship to tread element de
sign [14].

FIGURE 5. 16. Footprint region of a smooth truck tire;free rolling at I8m/s in 2.0mm of water
m-
CONTACT BETWEEN THE TIRE AND ROADWAY 261

FIGURE 5.17a Free rolling at 8.7m/'s.


Moire contouredfootprint of a belled bias ply tire [10]; water depth of 2.54 mm, inflation pres
sure of 130 kPa, height increment between fringes of 0.36mm. »J

Shape and size of contact areas for the same tire can differ considerably
when comparing on- and off-road operation. Figure 5.19 depicts both situ
ations. In general, on soft off-road surfaces the contact patch is larger and
shifted much further forward with respect to the axle position than on
paved surfaces [15]. Contact patch length (and width, depending on tire
construction) will be increased. The major factor causing these changes in
shape and size of the contact patch in off-road operation is the vertical
stress-strain (pressure-sinkage) relationship of the terrain material—the
softer the "soil", the larger and further forward will be the contact patch
[16].
5.2.2. Theory of Contact Area
No single unified theory exists for describing the area of contact of a tire
with a rigid roadway or test wheel. Most of the reasons for this lie in the
complexity of the tire construction, since from the point of view of mathe
matical modeling of the tire its shell-like structure and rather soft serrated
tread are difficult to represent analytically. Nevertheless, several special
ized theories are available for approximately describing the dimensions of
262 MECHANICS OF PNEUMATIC TIRES

FIGURE 5.17b. Free rolling at 13.4m/'s.


'ire contouredfootprint of a belted bias ply tire [1 0]; water depth of 2. 54 mm, inflation pres
sure of 130 kPa, height increment between fringes of 0.36mm.

the contact area of the tire with a roadway or test wheel, and these are use
ful in explaining some of the phenomena which are observed.
One type of problem which has been studied in some detail is that in
volving the contact of isotropic^olid elastic bodies with geometrically
simple surfaces. This solution obviously has application in various kinds
of bearings and friction drive devices. For the case of small deformations
of solid bodies in contact with one another, the contact areas have been
expressed in particularly useful form by Whittemore and Petrenko [19],
who conclude that
a. Contact area between an infinitely rigid plane and an elastic body of
double curvature, with principal radii of curvature Rc and Rt, as
shown in figure 5.21, is an ellipse.
b. Upon experiencing a deflection S against a rigid plane, as shown in
figure 5.22, the semi-major and semi-minor axis of the contact ellipse
are given by -—.
...- ^^
r — At»~l r OArtH
(5.1)

1
• \
CONTACT BETWEEN THE TIRE AND ROADWAY 263

FIGURE 5.17c. Free rolling at 18.1m/s.


Moire contouredfootprint of a belled bias ply lire (JOJ; water depth of 2.54 mm, inflation pres
sure of 130 kPa, height increment between fringes of 0.36mm.

-150

568
x
100 807

27.800
; 293,000
0.8 0.4 04 0.8 mm

FIGURE 5.18. Film thickness profiles for a rubber sphere sinking onto aflat glass plate with
squeeze times (in seconds) for each profile. [1 1] \ ~ 5461A.
264 MECHANICS OF PNEUMATIC TIRES

Direction of Trovel

Footprint
Length

FIGURE S.19a. Loaded, free rolling tire on a rigid surface.

where A is given by
"J
(5.2)

and a, /S and A must be taken from table 5.3 in terms of the angle θ, this
angle being denned by

cosff-
R, (5.3)

From these results it is seen that the isotropic contact theory, often
known as the Hertz contact theory, predicts linear dimensions which are
dependent upon the deflection to the exponent 0.5. Since the contact area
is an ellipse, it becomes linearly proportional to deflection in the form

i^} (5.4)

Direction of Travel

~*\ Footprint
Length

FIGURE S.19b. Loaded free rolling tire on a soft surface.


CONTACT BETWEEN THE TIRE AND ROADWAY 265

FIGURE 5.20. Photograph of contact patch of a loaded tire. (Courtesy The B. F. Goodrich
Tire Company.)

FIGURE 5.21. Radii of curvature K and R, of tire tread surface.


266 MECHANICS OF PNEUMATIC TIRES

FIGURE 5.22. Idealized contact geometry between a torus and a plane.

where the factor in parentheses is a function of the relative values of tire


radius Rc and tread radius R1.
While this type of solution certainly does not represent an automobile
tire completely, it is interesting to note that another situation does exist in
practice where the contact area can also be shown to be linearly depen
dent on deflection by quite a different form of analysis. This occurs in an
aircraft tire where the tread is relatively thin compared to automotive
tires. Here the cross section of the tire is close to being circular, with a
radius of curvature approximately equal to half of the section width w. In
this case, it has been demonstrated by many experiments that the contact
area between tire and roadway is also approximately an ellipse. Here it is
common to treat the tire as a carcass of zero bending rigidity, and of mem
brane characteristics such that it can map itself onto the flat roadway by
appropriate membrane compression of the carcass. A side view of this as
sumed tire geometry is shown in figure 5.23, where the tire with original
running band A OB is presumed to map itself onto the plane CC by assum
ing the shape A O′B. The geometry of this deformation is such that the
right triangle D&E is governed by the relation

(5.5)

TABLE 5.3.
10° 20° 30° 35° 40° 45° 50° 55°
a oo 6.612 3.778 2.731 2.397 2.136 1.926 1.754 1.611
ft 0 0.319 0.408 0.493 0.530 0.567 0.604 0.641 0.678
\ .851 1.220 1.453 1.550 1.637 0.709 1.772 1.828

60° 65° 70° 75" 80° 85° 90°


a 1.486 1.378 1.284 1.202 1.128 1.061 1.00
ft 0.717 0.759 0.802 0.846 0.893 0.944 1.00
A 1.875 1.912 1.944 1.967 1.985 1.996 2.00
CONTACT BETWEEN THE TIRE AND ROADWAY 267

DC=2RC

•' •

FIGURE 5.23. Tire contact symbols.

where it is assumed that the original and final tire shape is the same out
side of the contact area. From this one may obtain the contact length

(5.6)

If one considers the section of the tire at right angles to that shown in
figure 5.23, as illustrated in figure 5.24, then it is seen that an analogous
geometry governs the mapping of the tire in this cross-sectional dimen
sion. Using the same type of analysis as previously, the width of the con
tact ellipse b is given by the expression

(5.7),

FIGURE 5.24. Cross section view of a tire in contact.


268 MECHANICS OF PNEUMATIC TIRES

Since the contact area is elliptical in shape, its value is given by


i
A = "22
lb
"f (5.8)

Thus, for an aircraft tire, one might also anticipate a linear, or nearly lin
ear, relationship of area with deflection if the membrane carcass theory is
valid. Experiments with aircraft tires on this point have been conducted by
Michael [2] as quoted by Hadekel [1]. This data appears to be extremely
linear up to quite large deflections, and to be essentially independent of
inflation pressure, so that there appears to be some physical basis for ac
ceptance of the membrane theory for aircraft tires. Very similar relations
have been obtained from the data of Smiley and Home [18], with slight
modification in the numerical constants associated with the linear rela
tionship.
For an automotive tire the presence of the shoulder region causes the
contact area to be close to a rectangle whose length varies with deflection
of the tire. Two different methods have been proposed for calculation of
the contact patch length. The simpler one, which is useful mainly for static
or slow rolling considerations, utilizes an experimentally determined frac
tion of the geometric contact length given by eq (5.6). Experimental data
is given by Smiley and Home [18] for this using aircraft tires in figures
5.18 and 5.19, along with a curve showing the best fit obtained by using a ij
fraction of the geometric intersection length, in this case 85 percent of it. -4
A more complicated theory has been proposed by Bohm [20], Fiala [21],
and Clark [22] which attempts to take into account the dynamic character
istics of the contact patch. This model visualiy.es the running band of the
tire as consisting of a circular shell on an elastic foundation, the elastic
foundation representing the effect of sidewall support. In reference [22]
mathematical techniques are given for transforming these shell equations

— /M>c- ass Nnvi


26«66-IZPR-l -R23-A
26x6.6-l2PR-Sn-R23-B
40II2-I4PR-SD-R24-A
4O.I2-I4PR-YE -R24-B
56» 16 -24 PR-HE -R22 -A
56KI6-24PR-JZD-R22-B
57«20-I6 PLY-I156-INCH1-R53
L_i_l _ji_L_j J . I
02 04 .06 .08 .10 .12 .14 .16
VERTICAL-DEFLECTION PARAMETER, 8/Dc

FIGURE 5.25. Variation offootprint-length parameter with vertical deflection parameter for
several types I and Vll tires (aircraft).
CONTACT BETWEEN THE TIRE AND ROADWAY 269

//Oc-0.851(8/Dc-(8/Dc)

o 25x9 I Experimentol Dato _

.02 .04 .06 .08 .10 .12 .14


VERTICAL DIRECTION PARAMETER. 8/Dc

FIGURE 5.26. Variation offootprint-length parameter with vertical deflection parameter for
several German aircraft tires.

into moving coordinates so that the shell may be visualized as moving


I down a flat plane at some rolling velocity V,,. In both cases, it is possible to
calculate the length of contact patch by solution of the governing differen
tial equation.
On the assumption that the running band undergoes pure bending and
isjnextensible, one may solve the equations governing displacement of the
cylindrical shell on an elastic foundation for the specific case of contact
with a rigid frictionless plane, provided that one assumes that the initial
geometry of the shell continues to be valid. Under this assumption, it is
possible to assign values to the bending stiffness of the running band of the
tire, and to the elastic support given to the running band by the pressur
ized sidewalls. This allows direct numerical calculation of the contact
patch length as a function of vertical tire deflection. Such a typical calcu
lation is given in figure 5.27, compared with experiments taken from a tire
used as a model to obtain the stiffness and elastic spring support data.
Dodge [23] has compared the results of calculations and experimental tire
contact patch lengths for seven different automotive tires, five bias ply and
two radial. These comparisons are given in figure 5.28.
The foregoing discussion has concerned itself with gross contact area.
The presence of grooves and kerfs, designed to channel or wipe water will
substantially reduce this gross contact area so that the net contact area is
considerably less. The exact value of this must be determined for each in
dividual tread pattern. A representative set of different automotive tires
yields data as shown in table 5.4.
Fluid Film Effects
As has been determined experimentally, losses in traction that occur on
pavements covered by fluid contaminants result from a reduction in the
dry contact area. No comprehensive theory exists for predicting the frac
tion of dry contact area that will remain under all wet traction conditions.
However, analytical models have been proposed for estimating the extent
270 MECHANICS OF PNEUMATIC TIRES
14
Contact Potch Lengths
HOO- 14 Auto mo five 7. re
J3 '2 Comparison Calculation and
Tire Oiameter26.7m oprox.
1
T 10

VERTICAL TIRE DEFLECTION -


INCHES

FIGURE 5.27. Plot of contact patch length vs. vertical tire deflection.

and shape of the traction producing portion of the footprint under special
situations. Daughaday and Tung [24] mathematically modeled the sliding
of a smooth surfaced tire on a lightly wetted pavement. Only viscous
forces were considered in the water film. Their calculation procedure be
gins with the assumption of a deformed shape for the tire surface. This de
termines the extents and shapes of a thick film region and a semi-dry con
tact region. Traction forces are developable in the latter region. This film
shape is used along with the constraint of a specified imposed load to solve
a two-dimensional Reynolds equation for the sliding velocity. Thus, in this
(inverse) manner one can relate contact region with sliding velocity under
thin film conditions. Okamura [25] has proposed a more direct scheme for
predicting the extent and form of the dry contact area under similar oper
ating conditions. Experimentally determined influence coefficients are
used to determine tire deformation as part of an iterative scheme used to
obtain compatible fluid pressure (by solving Reynolds equation in the
fluid film) and film thickness distributions. Figure 5.29 shows the change
with speed of the dry contact area predicted by this formulation for a
smooth surfaced bias ply tire. Tire travel is from left to right. The numbers
next to the dotted lines denote the velocity in km/ h at which these lines
are the forward boundaries of the dry contact region.
The initiation of contact between a flexible tread element and a smooth
microtextured, or macrotextured surface has recently been modeled
mathematically by Browne, Whicker, and Rohde. [13, 14, 26]. Initial dry
contact is predicted to occur at the corners of rectangular elements and at
the outer edges of circular elements of the tread blocks, in agreement with
the previously reported experimental data. Figure 5.30 shows this charac
teristic corner type of contact.
CONTACT BETWEEN THE TIRE AND ROADWAY 271

I-T
Tl RE * " TIRE *z i
8: 10 1 14 EHAS PLY
12 < 12 7:50x 14 B AS PLY
I 1 o^S*
10 10
X
*>•' & ~sS?
8 8
6 X" 6
Jj^*' !
<F
'\°'/V \
o 4
X 4
X
2 2
* CALCULATED
O n
100 1.50 2.00
DEFLECTION- INCHES DEFLECTION -INCHES
12 TIRE *3 ~ —**•' TIRE
' i
10 5:90x15 BIAS

8
6
4

I 0 .25 50 1.00
DEFLECTION -INCHES
1.50 2.00V"0 .25 .50 1.00
DEFLECTION -INCHES
1.50 200

PATCH
LENGTH-lf 6
F
D
rat
&<r>o M *t »*5
.15! Smm «I5
T EXT LE
RAD AL
)
-^ EL!
^n
XJ! i

i i i
200
DEFLECTION -INCHES

FIGURE 5.28. Contact patch length-deflection characteristics for five automotive tires.

53 Normal Pressure Distributions


5.3.1 Experimental Measurements
Influence of Tire Variables
At the interface between the tire and the roadway an element of tire sur
face area is acted upon by a force vector which can be expressed as two
components, one perpendicular to the contact surface, called the normal
component, and one tangential to the contact surface. This is shown in fig
ure 5.1. This latter component may be further decomposed into two com
ponents, each lying in the contact plane, but one parallel to the central
plane of the tire and the other perpendicular to it. These components in
272 MECHANICS OF PNEUMATIC TIRES

10

o
:i ^ *°. "!
OO W> 00

Soo m
r^ oo
zlo o

v> m o Q <N

00

S
zlo
i*"l vo f^ O ^

II
V — e»i OO (N
m f^ <S <S H
0

•n

1 1111 1

oooo o

•» *n ^
1
-3-
x x x x H3 x
r-' oo »rl r* —
CONTACT BETWEEN THE TIRE AND ROADWAY 273

0 1 2 3 4 3 6 7 8 9 1 0~ I M2 13 14 ' IS 16 17 t8 19 20 21 22 c

FIGURE 5.29. Dry contact regions for different vehicle velocities [25J water depth - 1 cm;
5.60-13 4 pr smooth surfaced tire; inflation pressure — 1.4 Kg/cm3.

the contact plane are commonly called the shear components. With equal
validity they could be expressed as components parallel and perpendicular
to the direction of travel of the wheel. Either decomposition would be use
ful for describing the shear effects.
In this section attention is directed first to the normal pressure distribu
tion components caused by contact of the tire and some other surface. As
a basic primary concept, one might state that
P~Po + f (Tire structural characteristics, tire driving or braking torque,
tire side forces, tire velocity, etc.) (5.9)
where p is the vertical pressure component at any point, p0 is the inflation
pressure of the tire and / is some general functional relationship which in
sofar as is now known is extremely complicated, and can best be described
in a qualitative sense.
In eq (5.9), we postulate that the net pressure distribution at any point
depends primarily upon the inflation pressure, and there is considerable
experimental evidence to indicate that this is indeed true. Such a con
clusion was used by Hadekel [ 1] as the basis for one of his estimates of the
net contact pressure of a tire. On the other hand, subsequent studies have
shown that the characteristics determining the function / also play an im

Contoct ot Corners,

0.008cm

FIGURE S.30. Fluidfilm thickness distribution h(X, Z, T) under flexible tread element [13].
Load - 126N. Youngs modulus - 2580 Kl'a
T - 0.73 X 10 V contour increment - 0.07 mm.
274 MECHANICS OF PNEUMATIC TIRES

portant part in defining the detailed tire pressure distribution. In general,


these characteristics may be thought of as being divisible into two major
parts:
a. Tire structural characteristics.
b. Tire operating variables.
Modern instrumentation techniques have allowed measurement of nor
mal pressures under a variety of conditions, using many different types of
tires. Some of this data has been reviewed briefly by Hadekel, with partic
ular emphasis on the early work by Kraft [28], and by Markwick and
Starks [29] who dealt with automotive tires. More recently, much more de
tailed measurements on truck tires have been reported by Bode [30], who
reports data such as shown in figures 5.3 1-5.35. Additional data has re
cently been reported by Lippmann and Ohlizajck [31] on passenger car
tires. One set of contact pressure distributions taken from their work is il
lustrated in Fig. 5.36. A further comparitive study between various types
of tires has been reported by Freitag and Green [43] which is in part sum
marized in Figure 5.37 where pressure distributions taken both along the
length and width of three common types of automotive tires are presented.
These tires were tested in contact with an essentially rigid surface. They
illustrate the rather interesting point that the presence of a four-ply struc
ture with its greater bending rigidity tends to emphasize the pressure
spikes located at forward and leading edges of the contact patch, as might
be expected from the simple theoretical discussion of the bending of a
curved ring, given in Sec. 5.3.2. A similar but less detailed study has been
presented by Hofelt [44] as shown in figure 5.38.
Tires with tread patterns using many kerfs, or tires with small tread pat
tern blocks, exhibit complex pressure distributions. There are often large
variations in pressure over a small block. These are difficult to measure
since the size of the required instrumentation becomes small, and for this
reason only limited information is available on the effects of tread pattern
design on pressure distribution. However, it is known that the normal
pressure distribution between tire and pavement surfaces differs markedly

Up/cm2)
PRESSURE

FIGURE S.31. Pressure distribution under a lire with no tread pattern, i.e.. smooth surface.
Slant load 3740 Kp.. inflilum 6.5 aim gage, velocity 18.3 Km/hr
CONTACT BETWEEN THE TIRE AND ROADWAY 273

m c
V 8.
u 4
3 6
\ J \ J
\
ISo: 10° ,

FIGURE S.32. Pressure distribution down the length ofa tire with no tread pattern, at constant
IS Km/hr velocity, for two loads: (a) static load 1680 Kp., inflation 6.5 atm; (b) static load
3740 Kp., inflation 6.5 atm.

"100 60 60 40 20 0 20 40 60 80 100 mm
POSITION ACROSS WIDTH

FIGURE S.33. Influence of load and internal pressure on contact pressure distribution under a
tire with no tread pattern.
Constant velocity IS Km/hr measured across the tire width, (a) Load 1680 Kp, inflation 3.5 aim; (b) load 1680 Kp. infla
tion 5.0 atm; (c) load 1680 Kp. inflation 6.5 atm; (d) load 3740 Kp, inflation 6.S aim.

LEADING EDGE TRAILING EDGE

U
\
4 i\ c 1I
^~b
6 \- •\~~

in
"0 50 100 150 200 250 300mm
POSITION IN LENGTH OF CONTACT PATCH

FIGURE 5.34. Influence of the tractive force on pressure distribution under a smooth tire.
taken down the length of the tire.
Load 1680 Kp; inflation 6.3 aim (a) 0.6 m/sj acceleration V - 10.2 Km/hr, (b) 1.6 m/f2 acceleration V - 8.4 Km/hi, (c)
4.1 m/>2 acceleration V - 8.9 Km/hr.
276 MECHANICS OF PNEUMATIC TIRES

LEADING EDGE TRAILING EDGE


« 0

i2
* 4
u
? 6

I *
10,'0 50 100 150 200 250 3OOmm
POSITION IN LENGTH OF CONTACT PATCH

FIGURE 5.35. Influence of the braking force on pressure distribution under a smooth tire,
taken down the length of the tire.
Load 1680 Kp. inflation 6.5 aim. (a)- -4.1 m/s2 deceleration V - 10.0 Km/hr; (b)—2.7 m/V
Load 1680 KK inflation 6.5 aim. (a)—4.1 m/i2 deceleration V - 10.0 Km/hr. (b)—2.7 m/s1 deceleration V - 35.7 Km/
hr, it) -2 n m/s- decelention V - 14.5 Km/hi.

between smooth and patterned surfaced tires and also between tires with
different tread patterns. Working with otherwise similar tires, Tret'yakpv
and Novopol'skii [32] recorded normal pressure distributions, shown in
figure 5.39, for three different tread pattern designs—smooth surfaced,
block elements 3.5 cm by 3.5 cm with 20 percent void area, and block ele
ments 2. 1 cm by 2. 1 cm with 20 percent void area. Pressure levels in the
blocked patterns exceeded those in the smooth tread case by approxi
mately 25 percent. This can be equated with the fact that footprint dimen
sions remained nearly constant while void area increased 20 percent when
going to the patterned treads. Maximum pressure points occurred in the
middle of the projections, being slightly shifted to the "toe" of the tread
blocks, in the direction of vehicle travel, due to the increased stiffness of
the sheared and compressed rubber in that region. In an earlier study,
Markwick and Starks [29] found that uneven wear and moulding irregu
larities, including seams, could cause large distortions in the pressure dis
tribution and result in quite different distributions at different points in the

Axle centered
over tronsducer

FIGURE 5.36. Interfacial pressures of bias-belted H size; 26 psi, 0 deg camber and steer,
100% TandR load - 1580 Ib.
CONTACT BETWEEN THE TIRE AND ROADWAY 277

t--

l"

». 6.00-16. 4 -PR (STD CONST) SMOOTH TIRE

b. 8.00-16. 4 -PR (RADIAL PLY) SMOOTH TIRE

c. 16.00- 1& 2 -PR SMOOTH TIRE

FIGURE 5.31. Vertical stresses along principal axes of contact ellipse of three tires of
dissimilar construction under constant 890 Ib. load but with different inflation pressures.
278 MECHANICS OF PNEUMATIC TIRES

B. 2t pn inflation 10»» Ibi. C. 12 p

FIGURE 5.38. £^ecf of inflation and load on normal contact pressure.


(a) Effect of inflation it constant load, (b) effect of load at constant inflation pressure.

life of a tire. They also noted that for bias ply tires pressure varied in the
lateral direction with the peak occurring near the center as shown in Fig
ure 5.40. This was also recorded in two recent studies [31], [32].
There seems to be no record in the literature of studies of the effect of
tread compound on pressure distribution.

Influence of the Environment


The presence of fluid contaminants on the pavement surface can signifi
cantly alter the normal pressure distribution in the footprint of both
smooth surfaced and patterned tires. As a vehicle travels along a contami
nated pavement surface, its tires plow through the fluid and act to squeeze
it out of potential tread-pavement contact area. This causes pressures to
develop on the tread surface due to the inertia and viscosity of the fluid.
Increases in vehicle speed produce increases in pressure due to both fluid
properties. Portions of the tire surface on which these pressures equal (or
exceed) those found under dry contact conditions will be supported on a
fluid film. In fact, pressures exceeding those that would exist under dry
conditions can produce upward deformation of the tire surface allowing
thick pockets of fluid to persist. Figure 5.41 contains the fluid film thick
ness distribution found experimentally to exist under a smooth-surfaced
belted-bias tire. Maximum fluid pocket depth is 2.78 mm, nearly the thick
ness of the upstream water film. If sufficient fluid exists to flood tread
grooves, these portions of the tire surface will be pressurized. This effect is
shown in the experimental data plotted in Fig. 5.42. Such pressurization of
the groove regions (coupled with the fluid pressure-induced deformation
of the tire surface) can actually cause a lowering of under-rib pressures be
low dry contact levels. This is seen in the trailing portions of the pressure
profiles in Fig. 5.42 and over the entire length of the higher speed profiles
in Fig. 5.43.
The presence of fluid contaminants can also alter the pressure distribu
tions under individual elements of the tread pattern. The variation be-
CONTACT BETWEEN THE TIRE AND ROADWAY 279

80mm

80

PATTERNED TIRE
21 mm x 21 mm Trend Blocks

FIGURE 5.39a. Tread pattern effects on normal pressure distribution 6.00-16 tires, load — 460
Kg], inflation pressure — 2.2 Kg] /cm1, velocity = 1 km/h [32].

tween elements in pressure profile previously displayed in Fig. 5.39 dis


appears on fluid coated smooth surfaces, the pressure distribution under
all square elements there taking the general form shown in Fig. 5.44. The
nearly uniform pressure level is due to the pocket of fluid persisting be
tween tread element and pavement surfaces.
The asperities that exist on real pavement surfaces produce extremely
large though highly localized variations in dry contact interfacial pressure
levels. Giles and Sabey [34] have measured pressures of up to 8000 psi be
tween tire and pavement on the tips of sharp asperities. Similar high pres
sure intensities were recorded by Marwick and Starks [29] on the sharp
edges of stones in the pavement surface. They determined that
(1) For asperities of the same base width and height, the mean pressure
level between tire and asperity depends on the shape of the asperity,
sharper asperities having higher mean and peak pressures, and

'80mm

PATTERNED TIRE
35mm x 35mm Treod Blocks

FIGURE 5.39b. Tread pattern effects on normal pressure distribution 6.00-16 tires, load - 460
Kgf, inflation pressure — 2.2 Kgf/cm2, velocity — 1 km/h [32].
280 MECHANICS OF PNEUMATIC TIRES

80mm

SMOOTH SURFACED TIRE

FIGURE S.39c. Tread pattern effects on normal pressure distribution 6.00-16 tires, load - 460
Kgf, inflation pressure — 2.2 Kgf/cm1, velocity — I km/h [32J.

-3-2-101 2 3
POSITION ON MINOR AXIS, inches

Typicol print of contoct-el lipse of 8" x 24" ti re under


normal load and inflation pressure

FIGURE 5.40. Footprint and pressure distribution along the lateral centfrline for an 8" X 24"
tire, axle load — 2 tons, inflation pressure - 105 psi, speed - 40 mph [29].
CONTACT BETWEEN THE TIRE AND ROADWAY 281

TIRE-
PAVEMENT
CLEARANCE
(mm)

MX)
200
160
120 120
SIDE 80 LEAD
EDGE 40 EDGE
(mm) (mm)

FIGURE 5.4 1 . Fluidfilm thickness distribution under a smooth-surfaced hydro-planing tire (V


- I8.8m's, P* - 130 kPa).

(2) The mean pressure on an asperity is directly proportional to tread


rubber hardness.
These results can be summarized with the following equation:
E d
p=A (5.10)
— a2 a

where p is the mean pressure between asperity and tire tread, A isa con
stant depending on the sharpness of the asperity, E is the modulus of elas
ticity and o is Poisson's ratio for the tire-tread material, and a and d are
the base width and height, respectively, of the asperity. Figure 5.45 shows

Groove

Pressure
0.24 0.36 048 Gauge
RUNWAY DISTANCE, m Location

FIGURE S.42. Typical fluid ground pressure signatures obtained during passage of 32 X 8.8
type Vll aircraft tire over runway pressure plate. Tire pressure — ISO kPa; water depth — 2.5
cm. (33].
282 MECHANICS OF PNEUMATIC TIRES
Inflation: 24psi
Load : 1520 Ib
Tire: Bios Belt
Smooth with straight grooves
Free rolling tire in 1/2 in. of water

Center Rib
50
40mph
40
20
/
r ^1 /\
r^ \ /
SOmph
Jr\
ZJ X
60mph

FIGURE 5.43. Pressure on flooded road surface [1 7].

four different two-dimensional asperities for which the pressure distribu


tion has been calculated and the factor A determined. Solutions for certain
three-dimensional shapes including spheres and cones have also been re
ported in the literature [37].
A very complete set of data on normal stresses taken in various soils has
been reported by Green [41]. A typical pressure contour plot of a tire in
sand is shown in figure 5.46 while the total soil pressure distribution is
given more accurately by figures 5.47-5.49.
5.3.2 Theory of Normal Pressure Distributions
The use of theory is most beneficial in discussing the role of structural
parameters in modifying the basic inflation pressure distribution. Some of
the most important aspects of the various simple theories will be reviewed
here.
In the first instance, consider the contact against a flat surface of an in
flated membrane with vanishing bending stiffness. No matter what the
tension in the membrane, the fact that it is in geometric contact with the

1.9'

FIGURE 5.44. Fluid film pressure distribution P(X. Z, T) under flexible element [13].
Load - 126 N. Youngs modulus - 2580 kPa, T - 0.75 X ID'S.
CONTACT BETWEEN THE TIRE AND ROADWAY 283

Distribution
of Pressure

Form of
Asperity

Meon pressure , E i_
ponosperity TM? o

Uniform Distribution Distribution Distribution


Distribution of Pressure of Pressure of Pressure
of Pressure over Parabolic proportional over Wedge-
on Asperity Asperity to (oz-x2)8" Shaped Asperity

FIGURE 5.45. Two-dimensional surface asperities with calculated pressure-distributions [34].

flat surface means that the contact pressure distribution is exactly equal to
the inflation pressure inside the membrane, and it is this line of reasoning
which allows one to postulate that the primary component of tire vertical
contact pressures is the inflation pressure. On the other hand, if the tire is
in contact with a cylindrical road wheel then this statement is no longer
true, since the cover tension now plays some role in denning the net con-

DISTRIBUTION OF NORMAL STRESSES


PROJECTED ON HORIZONTAL PLANE
TOWED WHEEL IN SAND
1100-20, 12-PR SMOOTH TIRE
3OOO-LB LOAD IS-PSI INFL PRESS
0- TO 8-IN. CONE INDEX = 30

FlOURE 5.46. Distribution of normal stresses under a towed tire in sand. [42].
2*4 MECHANICS OF PNEUMATIC TIRES

EM AT AXLE

.\ T »(«)(frp
b

ASSUMPTIONS'

AXLE FRICTION NEGLIGIBLE

NORMAL AND TANGENTIAL


FORCES ON A TOWED
PNEUMATIC TIRE
IN SOFT SOILS

FIGURE 5.47. Stresses in a towed wheel in soft soil.

tact pressure. This is expressed analytically as

(5.11)

where N represents the shell membrane force per unit length while R rep
resents the radius of the road wheel contacting the tire. From this it may
he seen that the road wheel test may generally be expected to exhibit
slightly higher average contact pressures than actually exist on the high
way.
It is of interest to attempt to classify the various structural parameters of
CONTACT BETWEEN THE TIRE AND ROADWAY 219

HORIZONTAL COMPONENTS
•• TO • ->» C

VERTICAL
NORMAL STRESSES AT TIRE
SURFACE IN SAND
TOWED WHEELS
II 00-20. 12-PR SMOOTH TIRE

FIGURE 5.48. Normal stresses at tire surface in sand.

'tr»ta •• I n • •

-I 1 1 1 L

> Of T»,rf. ^

HOM MAt ( Mfk


HORIZONTAL COMPONENTS

NORMAL STRESSES AT TIRE


SURFACE IN SAND
TOWED AND POWERED WHEELS
1100-20, 12-PR SMOOTH TIRE
<6-PSI INFLATION PrUESJ = E

FIGURE 5.49. Normal stresses at tire surface in sand.


286 MECHANICS OF PNEUMATIC TIRES

the tire which affect the vertical contact pressure, in order to obtain some
idea of the influence of these parameters.
First of all, the tire is basically a structural shell. To an even greater ap
proximation, one may roughly consider a tire to be represented by an elas
tic running band elastically supported by the tire side walls. This type of
model is particularly applicable to a radial tire, although it also can be
used to express the characteristics of a bias ply tire. Using such a model,
attention may be primarily directed to the fore-and-aft variation of the
pressure distribution, since effects of right angles to this will be very simi
lar. This will allow description of the pressure distribution as a function of
only one dimension, namely that measured down the length of the contact
patch, which will considerably simplify the discussion which follows.
The tire structural components of most importance in describing the ef
fect of the tire carcass upon normal contact pressures are:
a. Elastic support of the tread by the sidewall.
b. Bending of the tread.
c. Shear deformation of the tread.
d. "Snap through" buckling of the tread, denned as the tendency of the
tread to seek a deformed equilibrium position due to membrane
compression.
e. Normal compliance or stiffness of the tread.
The role of these various components has been discussed from the theoret
ical point of view by Clark [44] who also performed model experiments
designed to illustrate some of these effects. If the tire is thought of as a
running band supported by an elastic foundation, then deformation of the
elastic foundation requires a pressure distribution which should be very
close to being proportional to the amount of radial deformation under
gone by the tire. The interference of the tire running against a flat plane is
known from the geometry, and the form of pressure distribution associ
ated with deformation of the side wall may be expressed as
j _ COS 00
(5-12)
COS0

where pa represents the contact pressure needed to deform the elastic side-
wall, k is the sidewall foundation stiffness in units of pressure, while fftl
represents the angle of geometric interference as shown in figure 5.23. This
type of pressure distribution is illustrated in figure 5.50. While in figure
5.51 an actual pressure recording down the length of the contact patch is
given by a model wheel made with an isotropic annular band. This is a
nearly pure elastic foundation and exhibits a pressure distribution similar
to that predicted by eq (5.12).
The bending of the running band itself represents an interesting form of
pressure distribution, since in considering the deformation of a circular
band in contact with a flat plane, a finite bending moment is required to
change the original curvature 1/r of the running band to a zero curvature,
which it must have when it conforms with the plane. In addition, this
CONTACT BETWEEN THE TIRE AND ROADWAY 287

FIGURE 5.50. Idealized elliptic pressure distribution.

change in curvature, and hence the bending moment, must be constant


within the region of contact. The only pressure distribution satisfying
these equilibrium and bending moment requirements is made up of a pair
of concentrated forces, located at each end of the contact patch. This is il
lustrated in figure 5.52 schematically, while figure 5.53 shows an actual
pressure distribution recording taken from a model test, in which a run
ning band free of elastic support is rolled over a pressure transducer bur
ied in a steel plate.
The case of pure bending of the running band is an idealization. In any
real situation some transverse shear deformation of the running band must
be present. The problem of the curved ring in contact with the flat plane,
utilizing both bending and shear deformation of the material, has been
studied by both Robbins [45] and Akasaka [46]. Both of these authors con
clude that the role of transverse shear deformation is to change the ideal
ized concentrated force distribution into a continuous one. The form of
this pressure distribution is illustrated in figure 5.54 in a schematic sense,
while a pressure recording taken from a worn automotive tire is given in
figure 5.55.
Buckling of the tread, or "running band" usually would occur at some
what larger deflections. This phenomena has been studied extensively
from the point of view of structural shell theory, although often using con
centrated forces or uniform pressure loadings in order to obtain stability
criteria. Experimental evidence gives a good idea of what happens here,

POSITION IN CONTACT PATCH

FIGURE 5.51. Measured pressure distribution of a silicone rubber ring with rigid hub.
288 MECHANICS OF PNEUMATIC TIRES

U U

/////////////////////////
FIGURE 5.52. Idealized pressure distribution for a circular ring.

although it is clear from the nature of the phenomena itself. Imagine an


internally inflated membrane subjected only to inflation and buckling ef
fects. The pressure distributions for small and large deflections respec
tively are shown in figure 5.56. Due to buckling, the pressure is decreased
markedly in the center of the contact region, as would be expected. Unfor
tunately the form of decrease of the contact pressure near the center of the
contact area is similar to that predicted by shear deformation, and it is dif
ficult to distinguish between the two causes of smaller contact pressure
near the center. Such a phenomena may also be observed on model tests by
increasing the deflection substantially, as is shown in figure 5.57.
The role of normal compliance, or stiffness of the tread, is not well un
derstood in forming the total pressure distribution pattern, since few per
tinent theoretical studies involving shell theory interacting with normal
stiffness exist. The general role of normal compliance can be seen by refer
ring to the well-known cylinder contact solutions of Hertz [47] for an iso-
tropic body. This approximates the type of contact pressure obtained
when a low normal stiffness exists, since most of the deformation due to
contact is confined to a relatively thin surface lamina. The Hertz solution
for a cylinder in contact gives a semi-elliptical pressure distribution as
shown in figure 5.50. Unfortunately, it is nearly identical in shape to the
pressure distribution required to deform the elastic foundation for the
tread, and it is difficult to distinguish between them. However, figure 5.57
shows a recording of the pressure distribution of a model having a rela
tively soft isotropic outer band simulating the tread of the tire. This illus
trates the general role of normal compliance, and illustrates further that
pressure distributions such as shown in figures 5.55 or 5.56 do not rise

POSITION IN CONTACT PATCH


FIGURE 5.53 Measured pressure distribution for a circular metal band on aflat plane.
CONTACT BETWEEN THE TIRE AND ROADWAY 289

FIGURE 5.54 Theoretical pressure distribution for a circular ring, with shear deformation, in
contact with a rigid plane.

abruptly, but rather exhibit some sort of gradual build up of pressure from
the edge of contact.
A number of theoretical investigations and mathematical models have
been proposed for explaining or calculating contact pressure distributions
between various types of bodies and a rigid plane. Hofferberth and Frank
[48] mention the work of Martin [49] as well as that of Vlassov [50]. In the
latter case a variational method is suggested for the treatment of two-di
mensional contact problems involving plates and shells, while in the
former case symmetric two-sided flattening of cylindrical and spherical

POSITION IN CONTACT PATCH

FIGURE 5.55 Measured vertical contact pressure on a 7.50-14 automotive tire with worn
tread.
290 MECHANICS OF PNEUMATIC TIRES

Lorgt Dtflfcthn

Small D»faction

POSITION IN CONTACT PATCH

FIGURE 5.56 Idealized pressure distributions due to buckling.

shells has been considered. Basically, the problem usually reduces to the
matching of two equations, one equation describing the free surface of
that portion of the plate or shell not in contact with the rigid plane, and
this defining equation usually is written in terms of known deformations
but unknown load distribution. Continuity of displacement must take
place and hence the two sets of equations must be matched, usually at the
unknown boundary of contact. This introduces an additional degree of
freedom into the system, and often makes such a matching process ame
nable only to trial and error solution. Bohm [51] has extended Martin's
theory and developed it for the one-sided flattening of the toroidal shell
with circular cross section, but without numerical results. Martin found
theoretically that the contact pressure distribution is rectangular under a
flattened cylinder, with a value smaller than that of the inflation pressure
providing the cylinder has bending stiffness. Equilibrium is maintained by
a pair of concentrated bending moments which act at the forward and
leading edges of the contact patch, these bending moments arising due to
the fact that it is not possible to match bending moment and shear at the
junction of the contact region with the free surface. On the other hand, the
inclusion of shear deformation by Akasaka and Robbins leads to theoreti-

POSITION IN CONTACT PATCH


FIGURE 5.57 Measured pressure distribution on tire model at small and large deflections.
CONTACT BETWEEN THE TIRE AND ROADWAY 291
2.0

5° 10° 15" 20° 25° 30*


POSITION IN CONTACT PATCH, &-

FIGURE 5.59 Calculated pressure distributions under a flattened cylindrical shell, under
internal pressure p, with different degrees offlattening.
The angles given are characteristic for half of the flattened, originally cylindrical portion of the arc.

cal pressures as shown in figure 5.58. An excellent summary of the current


status of contact mechanics is given in the recent review paper of Kalker
[156].
A great deal of interest has been shown in the role of velocity in modi
fying the contact pressure distributions just discussed. A number of inves
tigations have examined this, including the previously discussed work of
Bode as well as the work of Zakaharov and Novopol'skii [52]. In general,
the results seem to show that increasing speed causes an increasing vertical
contact pressure at the forward end of the contact patch and a decreasing
value at the rear portion of the contact patch. This is schematically illus
trated in figure 5.59, and has been experimentally confirmed by Zakaha
rov and Novopol'skii. Attempts have been made by Clark [22], to utilize a
relatively simple shell and elastic foundation theory to predict dynamic
contact pressures, but due to the inability of such a simplified model to ac
curately provide for continuity of shear and bending moment through the
edge of the contact patch, the corresponding dynamic pressure distribu
tions are predicted in the form shown in figure 5.60. It is clear that shear
deformation must be added to such a theory before realistic contact pres
sure distributions will be available. However, the general trends of such
dynamic pressure distributions are apparent.
The necessity for including shear deformation effects, and for dis
carding the restrictions of the Kirchoff bending hypothesis, were also dis-

Direction
of Travel

Pressure

FIGURE 5.59 Idealized pressure distribution as influenced by speed of travel.


292 MECHANICS OF PNEUMATIC TIRES

Direction
of Travel

Ola IHb
I - Inflation Pressure
n - Elastic Sidewall
Support
IHa - Forward Edge
Shear Disconuity
ffib - Trailing Edge
— Shear Disconuity

FIGURE 5.60 Calculated dynamic pressure distribution from simple elastic shell tire model,
without shear deformation.

cussed by Essenburg and Gulati [53] who studied the axisymmetric con
tact of plates. They concluded that only by such a relaxation of
conventional bending theory could the contact problem be successfully
approached.
All measurements on automotive tires show a relatively high normal
contact pressure region in the shoulder area of the tire tread, due to the
heavy tread shoulder conventionally used on such tires. This means that
generally one finds higher pressures at the shoulder than at the center of
contact. This has been illustrated in the experimental data given, but bears
repeating.
Interfacial pressure distributions for rubber loaded against pavement
asperities are a key element in many models of the sliding friction of rub
ber on both dry and lubricated textured surfaces. In such models, simple
empirical and analytical formulations, including educated guesses have
been used for the pressure distributions. Discussion of these will be de
ferred to the section on thin film wet traction. Recent efforts by Rohde [26]
and Browne and Whicker [14] have provided an analytical basis for deter
mining the combined effects of tread element flexibility, fluid con-
taminent, and pavement microtexture in two dimensions, or macrotexture
on interfacial normal pressure distribution.

5.4 TANGENTIAL PRESSURE DISTRIBUTIONS


5.4.1 Experimental Measurements
The tangential component of the contact stress between a tire and road
way has been studied rather extensively due to its close association with
the processes of tire wear and skid. Hadekel [1] quotes from the early work
of Martin [49], Mark wick and Starks [29] and Kraft [28]. Most of these
early measurements were made with mechanical dynamometer systems
imbedded in a flat plate, which allowed a rolling tire to actuate a recording
system and to measure the resulting tangential surface stress in terms of its
components in the direction of travel and at right angles to the direction of
travel In particular, Martin has idealized these measurements into two
vector components, each of which must be doubly symmetric about the
centerline of the contact area for a stationary tire without side or longitu
dinal force acting on it. This is illustrated in figure 5.61. These components
CONTACT BETWEEN THE TIRE AND ROADWAY 293

FIGURE 5.61 Shear stress distribution in the contact patch of an aircraft tire.

combine in such a way as to form shear stress resultants which are roughly
directed toward the center of the contact zone.
The distribution of tangential forces as given in figure 5.61 is idealized,
since there the tire is presumed to have been loaded statically and not
rolled into position. In this case it is clear that the deformation of the non-
developable tire surface, so as to cause contact with the flat roadway, will
almost surely result in both bending and membrane deformations of the
tire carcass with consequent symmetric strain distribution. The fodionson
efficient between the tread and roadway prevents the free contraction or
expansionUfthe tread surface, so that local shear stresses are set up in the
tread elements, between the essentially rigid roadway and the deforming or
straining tire carcass. It is these shear stresses which on the road surface
become the tangential force components during straight line rolling, and
at the same time serve to bring the elastic tire carcass into the state of com
patibility of deformation with the entire roadway and tire structure sys
tem. Were it ^ not for the coefficient of friction between the
and roadwayr^ejEe,.carcass would freely deform and there would be no
tangential stress distribution. Were it not for the elastic characteristics of
the tire tread elements and the tire carcass, there would be complete slip
throughout the contact region as deformation took place. In truth, the ac
tual phenomena lies somewhere between these two conditions. Under con
ditions of static deformation of the tire against a roadway, with appre
ciable friction coefficient, most of the contact surface is relatively free of
slip, while in cases of heavy braking or yawed rolling, substantial amounts
of primary slip are observed.
Detailed experiments on the friction of rubber indicate that, potentially,
friction coefficients are available which exceed those normally observed in
a typical pneumatic tire in contact with a roadway. It is felt that the cause
of this probably lies in the fact that small amounts of slip (secondary slip)
exist almost entirely throughout the contact patch for most practical cases
of tire rolling, so that in effect the measurement is of the coefficient of fric-
294 MECHANICS OF PNEUMATIC TIRES

LOAD 186 LB. ,


INFLATION PRESSURE: 40 LB/INZ
DIAMETER OF PLUNGER: 1/8 IN .
DEPTH OF PROJECTION: O.OI IN

ARROW DENOTES
DIRECTION OF
SHEAR STRESS

CENTER OF TYRE

-1.0 -0.5 0 0.5 1.0


POSITION ON MINOR AXIS OF CONTACT -ELLIPSE : INCHES

FIGURE 5.62 Shear stress distribution across the width of an aircraft tire.

tion of rubber in a state of secondary slip, and not in a state of maximum


friction coefficient.
The apparent symmetry of the static shear stress distribution is shown in
detail by some of the early experimental results of Martin, given in figure
5.62 for the shear stress component perpendicular to the midplane of the
wheel, i.e., in the lateral direction.
The exact distribution of the lateral component of the tangential contact
stressls not well understood even for a stationary tire. The work of Bider-
man, Volodina and Pugin [55] seems to indicate that for a standing or slow
rolling tire the lateral (sideward) component of the tangential contact
stress starts from zero at the forward edge of the contact area, reaches a
maximum some distance away from this point and then decreases to zero
at the trailing edge of the contact area. This is inferred from meridional
displacements of the carcass, but is in complete accord with figure 5.62.
Novopol'skii and Nepomnyashchii [54] have studied the longitudinal
component of the tangential stress vector in some detail, and have pro
posed two causes for this stress component in a rolling tire without steer
angle. These are:
a. Deformation of the carcass with respect to the contact plane, giv
ing rise to a shear stress approximated by the expression

T=k - sin 200 - sin 20 (5.13)

where T = longitudinal shear stress


k = carcass stiffness constant
and the other symbols are defined in figure 5.23.
b. A uniformly distributed shear stress caused by tire rolling losses,
which foT commercial automotivelires is a negligible factor compared l!
to carcass deformation. <=*
CONTACT BETWEEN THE TIRE AND ROADWAY 295

TRAILING
EDGE POSITION

£»<?£-
\
Position along contact length.

FIGURE 5.63 Distribution of longitudinal tangential stresses along the contact length between
a tire and the road while the wheel is in motion.

The form of the shear stress variation down the length of the contact patch
predicted by eq. 5.13 is given in figure 5.63.
Lippmann and Oblizajek [33] have recently published detailed mea
surements of tangential pressures in the tire longitudinal direction. Gener
ally these agree with the general conclusions just reached, although they
do show an abrupt reversal of the sign of the fore-aft tangential force at
the extreme trailing edge of contact. A typical record of this type is shown
in figure 5.64. This type of behaviour has also been reported by Seitz and
Hussman [56], although no clear explanation for it is at hand.

RIB NUMBER

Axle centered
over tronsducer

FIGURE 5.64 Fore and aft tractions of bias-belted H size; 26 pa, 0 deg camber and steer,
100% T and R load- 1580 Ib [31].
296 MECHANICS OF PNEUMATIC TIRES

Extensive experimental work on the longitudinal component of the con


tact shear stress has recently been published by Bode [30]. Here, an effort
is made to assess the role of traction and braking in forming the longitudi
nal component of tangential contact stress. Figure 5.65 shows a three-di
mensional plot of such a stress component over the entire contact area,
while figures 5.66 and 5.67 from Bode show that the influence of either
braking or tractive forces is to throw the major part of the longitudinal
component of tangential stress to the rear of the contact patch, as one
might anticipate. The direction of this component of the tangential stress
is dependent on the presence of tractive or braking forces. Even moderate
values of such forces are sufficient to completely change the distribution of
tangential stress, again as might be expected.
ItJs possible to separate out the effects of traction on shear from the
normal tire deformation effects on shear. This is done in figure 5.68, where
the heavy lines show the net shear distribution taken in the longitudinal
direction down the length of the contact patch, and caused directly by the
tractive effort. This is obtained by subtraction from the data of figure 5.66,
while a similar construction can be performed for the effects of braking on
shear. This is shown in figure 5.69, again taken from Bode. From this it is
seen that the effects of traction and braking are essentially mirror images
of one another insofar as they influence the magnitude and distribution of
the longitudinal component of tangential force in the contact area.
The experimental data given by Bode agrees with the measurements of
Novopol'skii and Nepomnyashchii.
The lateral component of the tangential stress, at right angles to the di
rection of tire travel has been extensively investigated by Gough [58], [57]
and by Cooper [59] for the case of yawed rolling, where the lateral tan -
gential forces are much larger than exist for straight-One rolling. In the

SHEAR

TO DIRECTION
OF TRAVEL _nl

SHEAR IN
DIRECTION
OF TRAVEL

FIGURE 5.65 Distribution of shear stress in the running direction of a smooth, braked tire.
CONTACT BETWEEN THE TIRE AND ROADWAY 297

Acceleration
0.96 m/s*
1.28 m/s*
1.53 m/s*
2.02 m/s*
3.22 m/s*
5.66 m /$*
Velocity
o: V-4l.7km/h
b: V« I 7.0 km/h
c: V = 25.8km/h
d: V= I7.7km/h
t: V= 7.4km/h
f: V = 8.9km/h
"0 50 100 150 200mm
LENGTH ALONG CONTACT

FIGURE 5.66. Distribution of shear in Ike running direction of a smooth tire under tractive
force.
Sunc load 1S90 Kp . inflation 6.S aim

case of yawed rolling, one may think of the tangential stress distribution as
first associated with forcing the elastic nondevelopable pneumatic tire
against a flat roadway, followed by the addition of effects due either to
yawed rolling, braking or acceleration. Viewed in this way, one might pos
tulate that the yawed rolling, braking or acceleration effects are clearly
dominant. Figure 5.70 taken from Cooper, shows the distribution of cor
nering force intensity as a function of tread lateral distortion at the center
of a yawed tire. Cornering force intensity is obtained by integrating the
lateral component of tangential stress across the width of the tire contact
patch. A more complete set of such curves is shown in figure 5.7 1 from the
same source. Figure 5.72 shows the cornering force intensity plotted
against position in the contact patch, giving a much clearer picture of the
asymmetric nature of this force component due to yawing. In figure 5.72,

deceleration Slip
SHEAR OPPOSING^ a: -0.4 m/s*
TRAVEL DIRECTION *b - 1.2%
b: -1.23 m/s* •b = 3.0%
C: - 1 .5 m/s* sb * 4.8 %
d: -2.45 m/s* •b = 6.9%
e: - 2.78 m/s* •b = 10.1%
f: -4.3 m/s* * 19.3%
Velocity
0: V= 17.6 km/h
b: V=l7.7km/h
c: V=I6.9 km/h
d: V* 15.4 km/h
•: V'16.9 km/h
f: V* 17.2 km/h

0 50 100 150 200 250mm


LENGTH ALONG CONTACT

FIGURE 5.67. Distribution of shear in the running direction of a smooth tire under braking
force.
Static load 2610 Kp., inflation 6.5 aim
2M MECHANICS OF PNEUMATIC TIRES

«, 50 SHEAR OPPOSING
DIRECTION OF MOTION '7C
.^

ji4.0
in
ffi 3.0
1
0,2.0
<
UJ
£ i.o
OX),
0 50 100 150 200 250mm
POSITION ALONG CONTACT LENGTH

FIGURE 5.68. Distribution of shear along the length in running direction.


Super-position of shear due to deformation and acceleration. Heavy lines are shear stresses due to acceleration alone.

the centroid of the area under the cornering force intensity curve lies be
hind the geometric center of the contact area, here denoted by 0, which
gives the self-aligning torque to the tire. The distance t in figure 5.72 is
called the pneumatic trail. A more complete set of curves of cornering
force intensity as a function of position in the contact patch are given for
various yaw angles in figure 5.73. Finally, the results of such measure
ments may be combined into a single plot such as shown in figure 5.74.
Due to the method of making these measurements, it is not possible to
determine the exact distribution of the lateral (sideward) component of
tangential contact stress across the width of the contact area. Only the in
tegrated value is given.
In figure 5.74 it is seen that the pneumatic trail decreases substantially
as the steer angle increases. This is related to the shape of the cornering
force intensity curve with increasing steer angle, since the greater slip in
the contact patch results in a more symmetric cornering force distribution.

™0 50 100 150 200 250mm


POSITION ALONG CONTACT LENGTH

FIGURE 5.69. Distribution of shear along the length in running direction.


Super-position of shear due lo deformation and braking, from figure 3.36. Heavy lines are due to braking alone.
CONTACT BETWEEN THE TIRE AND ROADWAY 299

(Ib-
;FORCE
IONRTNENRSINTGY
Tire Size: 6.70 -16
Pressure: 16 psi
^V
\ LoSl
S
8
S od:6-l cwt
p Angli»:4»

/
/ \
1 \

*
E
0.1 0.2 0.3
\
<
'*
0.4
Ul

SIDE SLIP (inches)

FIGURE 5.70. Cornering force intensity vs. tread lateral distortion.

The interaction of vertical pressure distribution with braking and yaw


has been studied experimentally by Iritani and Baba [60]. They show that
there is little change in vertical pressure due to yaw, but a measureable
amount due to braking. This latter effect is shown in figure 5.75.

5.4.2. Theory of Tangential Pressure Distributions


The action of a driving torque changes the magnitude and distribution
of the tangential forces over the plane of contact. No equation has yet
been found to connect the tangential force and the torque, but a qualita
tive picture can be drawn by considering the rotation of an elastic wheel.

100
Tire Size: 6.70-16
Pressure: 16 psi

O.I 0.2 0.3 0.4 0.5 0.6 0.7 OS 0.9 1.0 I.I 1.2
SIDE SLIP (inches)
FIGURE 5.71. Cornering force intensity vs. tread lateral distortion at slip angles up to 12°.
300 MECHANICS OF PNEUMATIC TIRES

TIRE SIZE: 6.70-16 SAT-CFxt


60 PRESSURE: I6psi CF-AREA
LOAD: 6-1 cwt t-TRAIl
SUP ANGLE: 4
o
cr ;
O
<9 = 20
£
UJ
Z
a:
O
O CONTACT POSITION (inches)
FIGURE 5.72. Cornering force intensity along the contact spot.

Figure 5.16 shows the rotation of a wheel transmitting torque A/,. As a


result of the torque transmitted through the wheel, two sets of forces act
upon it—the reaction of the wheel axis Pk, and equal to it in the opposite
direction the reaction ofthe road acting in the plane of contact. As a first
approximation it may be assumed that the reaction ofthe road Pk is evenly
distributed over the area of contact. The component of tangential stress
from the reaction Pk is denoted as τp (see fig. 5.76).
There is a certain amount of adhesion over the area of contact, and be
cause the tire possesses tangential elasticity the torque will compress the
tread elements in the zone immediately before contact (denoted by the
sign —) and at the same time will stretch those elements in the area just
after contact has been released (denoted by the sign +). Thus, an initially
compressed element, AJC, of the tread, is released from this compression as
it passes through the contact area, reverting to its normal state ∆x1, then it
undergoes stretching and emerges from contact in a stretched state ∆x2.
Since the elements, as they pass into the contact area, are in direct con-

Pressure: 16 psi
Load: 6- 1 cwt

-3 -2-1 0 12
CONTACT POSITION (inches)

FIGURE 5.73. Cornering force intensity at slip angles up to 12".


CONTACT BETWEEN THE TIRE AND ROADWAY Ml

O
Tire Size 6.70-16
Pressure: 16 psi
Load: 6.1
- 400

UJ
O
g300
C
o
£ 200

I
O

0 2 4 6 8 10 12 14
SLIP ANGLE (degrees)

FIGURE 5.74. Cornering force and trail at slip angles up to 12°.

BRAKING FORCE
•OKg
055 kg
x| 1 5kg
1 70 Kg

864202468 10
LEADING EDGE TRAILING EDGE
CONTACT LENGTH cm
TIRE 560- 14 4ply
LOAD • 300kg
INFLATION' I 55kg/cm«

FIGURE 5.75. Verticalforce and braking force distributions.


302 MECHANICS OF PNEUMATIC TIRES

FIGURE S.76. Rotation of a driving wheel (the distribution of longitudinal tangential stresses
in the contact zone of driven, driving, and braked wheels).
OXotal component of the longitudinal tangential stress rm due to the driving torque: (2) component r. due to the brak
ing torque; (3) distribution of longitudinal tangential stress along the contact length of a driving wheel; (4) ditto, for lo of a
free wheel; (5) ditto for a braked wheel.

tact with the road, any change in their dimensions is prevented by the
force with which each element grips the road surface, and tangential
stresses r* arise in the plane of contact. As a first approximation it is as
sumed that the stress τk changes linearly along the length of the contact
zone, at the center of contact τk = 0 (see fig. 5.76). Thus tangential stress
τm, which is the sum of τk and τc is also linear and is represented by
straight line 1 in figure 5.76.
When the wheel is braked the situation is reversed, i.e.. immediately be
fore contact the tread elements are stretched, and they contract at the
point at which contact is released. In this case it is line 2 (fig. 5.76) which
represents the tangential stress τm.
Thus, the magnitude and direction of the longitudinal tangential
stresses, τ, acting in the plane of contact of a driving or braked wheel are
determined by the sum of the stresses created in the rolling of a free wheel,
τ0, and the additional stresses τm created by the application of a torque.
The lines representing the resultant tangential stresses along the length of
the contact zone of a driving or braked wheel are shown at the bottom of
figure 5.76. The precise way in which these stresses are distributed over the
contact area depends on the design of the tire, the radial load, the internal
air pressure, and the adhesion to the road. These tangential stresses can
cause local slipping of the tread elements on the road. It is clear that
radial, bias and other tire constructions would exhibit different detailed
shear stress distributions, even for the same tractive force.
CONTACT BETWEEN THE TIRE AND ROADWAY 303

Vermeulen and Johnson [61] have discussed the contact of nonspherical


elastic bodies transmitting tangential forces, and have by such an effort
given at least a beginning to a formal theory for tractive forces in combi
nation with the contact problem. Vermeulen and Johnson presume that
the presence of tangential forces does not, to a first approximation, change
or alter the normal pressure distribution. Their discussion further pre
sumes that the contact surface of two rolling bodies transmitting a tan
gential force is divided unsym metrically into a region of slip and a region
of no slip. The no-slip region is adjacent to the leading edge of the contact
boundary, independent of the direction of the torque. Elliptical contact re
gions are assumed, and it is also assumed that regions of no slip are lo
cated adjacent to the contact ellipse at the leading edge.

5.5 Slip Between Tire and Roadway


In all of the discussion which has been given here, nothing has been said
of the role of slip in the contact patch. Since this becomes important for
purposes of describing the yawed roll i ng tire, as well as in braking and ac
celerating a vehicle, it is necessary to discuss this question briefly.
The term "developable surface" is generally used to denote a curved
surface which can be formed by simple folding or bending of a flat sheet,
without extension or compression of this sheet. A cylinder is an example
of a developable surface, as is a cone. The concept of a developable sur
face is important since it represents a kind of surface which may be placed
in finite area contact with a flat plane by simple bending processes without
stretching, tearing or compressing the sheet. Both the cylinder and cone
exhibit this kind of behavior.
A surface which is not developable can only be formed from a flat sheet
by means of stretching or contracting the sheet in its plane. A toroid,
which is generally the shape of most tires, is not developable. Thus, when
a tire is brought into contact with a flat roadway such a geometric process
can only take place by both simultaneous bending and compressing of the
tire surface. This means that tread elements in contact with the roadway
will generally undergo a small deformation in the plane of the carcass as
such tread elements pass through the contact patch of the rolling tire and
exit out the trailing edge. This process takes place in every tire. It is pos
sible to minimize the membrane stretching and contraction by making the
carcass structure of the tire as rigid as possible. This is generaUyJhe prin
cipal behind the radial or belted tire constructions. It is also possible to
minimize such membrane distortions by use of extremely stiff materials
foTThfTfire: carcass, such as steel wire or glass. Since the membrane defor
mations actually take place in the carcass of the tire, it should be possible
to prevent motion of the tread on the contact surface by choice of an ap
propriately soft tread compound. Unfortunately this normally counteracts
attempts to secure a long wearing tread material, so that it is usually not
possible to make the shear stiffness of the tread itself low enough to pre
vent all motion. This description just given applies primarily to straight
line motion of the tire, and describes what might be thought of as second
ary slip between tread and roadway, that is to say, slip whose magnitude is
of the order of a few thousandths of an inch while a panicle on the surface
of the tire moves through the contact patch. The direction of such second
ary slip may be partly longitudinal and partly transverse, depending on
tire construction and local friction conditions.
304 MECHANICS OF PNEUMATIC TIRES

Over and above secondary slip exists a phenomena which might be clas
sified as primary slip. This normally takes place independent of the non-
developable nature of the toroidal tire, and can be thought of as a function
of such parameters as steer angle, braking torque, and driving torque act
ing on the tire. The most common of these phenomena is that of steer
angle. Here, the contact area of the tire with the roadway is first of all dis
torted due to the presence of the steering effort, and finally is separated
into two regions due to this steering effort. The first region is one of essen
tially static contact between the tire and the road surface, but nevertheless
a region in which secondary slip may exist. The second region, adjoining
the first, is the region of primary slip where the tangential surface stresses
necessary to maintain geometric distortion of the tread surface exceed the
local frictional stresses available, so that the tire surface exists in part in a
state of slipping in a gross sense. The order of magnitude of such primary
slip might be in the neighborhood of 0. 1 inches or more.
It is difficult to measure the contact areas under such conditions of
yawed rolling, but these have been observed on model tires through a
glass plate by Saito [62]. Drawings are given in figure 5.77 showing the
growth of the primary slip area with increasing steer angle.
It should be noted in figure 5.77 that the primary slip is confined to the
rear of the contact patch in all cases. Effective stationary contact between
tire and roadway, marred only by secondary slip, occurs always at the
leading edge of the contact patch as the tire rolls.
In regard to straight line motion of the tire under braking or tractive ef
fort, there is probably always a certain amount of primary slip when the
wheel is in this condition, but the magnitude of this slip is variable. The
size of the zones of slip may also change, depending on the magnitude and
direction of the applied wheel torque. At the point where a driven wheel
comes into contact with the road and also at the point where it leaves the
road, the tire surface attempts to slip in the direction in which the wheel is
turning, while at the center of the contact area there is a region where the
tendency is to slip in the opposite direction. Novopol'skii and Nepom-
nyashchii [55] present diagrams showing the change in the positions of
tread elements along the contact length of a tire relative to the positions
which they occupied at the instant of entering contact. This is shown in
figure 5.78. The supping of a driving or braking wheel is also determined

A - adhesive area (secondary slip area)


B - partial skidding area (primary slip area)

FIGURE S.77. Slip areas as a function of steer angle <j> as determined by visual observation.
CONTACT BETWEEN THE TIRE AND ROADWAY 305

FIGURE 5.78. Displacement of tread elements along contact length of tire: (a) free rolling, (b)
driving, (c) braked.
Zone I Longitudinal tangential stress acting from tire to roadway in direction of motion.
Zone II Longitudinal tangential stress acting from tire to roadway opposite to direction of motion.

by the magnitude of the tangential forces acting in the plane of contact.


Figure 5.78 also shows displacement curves for the tread of a driving
wheel and for the tread of a braked wheel. Note that in both of these latter
cases the tendency for primary slip is confined to the rear of the contact
patch, just as in yawed rolling shown in figure 5.71.
This type of behavior is exemplied by measurements taken on an auto
motive tire and presented in reference [54]. This data is given in figure
5.79 where it is shown that the slip measured at the surface of a tread ele
ment is essentially in accord with what one would expect from the ideal
ized representation of figure 5.78.
In these cases the differentiation between primary and secondary slip re
gions is not particularly clear. For example, the first plot of figure 5.79
shows a freely rolling wheel without applied torque. Here, essentially all
of the tire is at a slip level low enough so as to be classed as secondary slip.
On the other hand, the last plot of figure 5.79 shows the slip of a braked
wheel in which slip values are large enough so that they might be thought
of as primary slip toward the end of the contact region. It is clear, of
course, that in the case of steering under fairly large steer angles the slip at
the rear of the contact patch might be primary slip, of the order of cen
timeters, while under complete locked wheel skidding the slip values
would be very large, and the entire contact patch would be in a state of
primary slip.
One comment should be made concerning the measurements of slip
commonly available in the literature. They are obtained in almost all cases
by rolling the tire over a plate which has protruding from its smooth sur
face a small pin of low mechanical impedance. The pin contacts the tread
as the tread element moves into contact with the roadway, and from that
point onward registers a signal proportional to slip as the tread element
moves through the contact patch. Thus all measurements are made with
respect to the leading edge of the contact patch, which serves as a zero ref-
306 MECHANICS OF PNEUMATIC TIRES

100 75 50 25 0 25 50 75 100 mn

FIGURE 5.19. Longitudinal slipping in the contact zone of a 6.00-16 tire.


(a) driven wheel-zero torque, (b) driving wheel-torque - 25Kgf • m . (c) braked wheel-torque - —25Kgf • m.
Contact length measured along the abscissa, with A the point of first contact and B the point of last contact. Ordinate is
longitudinal displacement of a tread element from its initial position. Downward is in the direction of turning of the wheel.
Curve 1, middle of tread; Curve 2, 17mm. from tread centerline; Curve 3, 34mm. from tread centerline; Curve 4.51mm.
from the tread cenlerline, i.e.. edge of tread.

erence. The data previously quoted have pertained to slip in the longitudi
nal direction, i.e., slip in the direction of travel of the rolling wheel. Bider-
man, Volodina, and Pugin [55] have used similar techniques to measure
deformation of the tire carcass. While this is not the same as the slip be
tween tread elements and the roadway, it nevertheless takes on the same
character. Transverse slip, at right angles to the direction of travel of the
wheel, also takes place in the contact patch. Its nature may be inferred
from figure 5.80, which gives data on meridianal displacement of the tire
carcass itself as the tire rolls through the contact area.
CONTACT BETWEEN THE TIRE AND ROADWAY 307

80 40 0 40 80 120
LENGTH ALONG CONTACT AREA, mm

FIGURE 5.80 Meridional displacement of carcass.


Distance from crown: I Mlmm. 2 2mm 3-40mm; 4-30mm; 5 55mm Straight line stow rolling. 6.00-16 tire, longitudinal
ribs 2mm high

It is clear that the question of slip during the rolling process is in


timately connected with the normal pressure distribution which exists in
the contact area. The normal pressure distribution does not suddenly rise
at the edge of the contact area from zero to a finite value, but rather in
creases, however gradually or suddenly, at a finite rate. This means that
inevitably there will be a narrow ring or boundary layer around the edge
of the contact patch on which the normal pressure distribution is very
small, so that consequently the available adhesion limits are very small. At
these regions it would be natural to expect that at least secondary slip
would occur simply due to the fact that horizontal friction forces are un
able to further restrain elastic displacement of the rolling tire. The experi
ments of Mindlm [63] tend to substantiate this.
An analogy to this has been given by Reynolds, and is quoted by Hade-
kel [1]. Imagine a belt on a pulley, transmitting a definite power, and
therefore having a difference in tension between entry and exit, and hence
also a difference of strain between these points. A given element of the belt
undergoes some change in length from entry to exit, which is possible only
if there is relative sliding. Reynolds concluded that there must be some
sliding near the edges of the contact region, even in the absence of hori
zontal forces other than rolling resistance.
The question of slip in a rolling tire has been studied by Novopol'skii
and Tret'yakov [64] who conclude:
a. When a tire rolls, the slip areas are where contact begins and ends.
b. Torque greatly increases both the area of the slip and the amount of
slip where contact ends. There is only an insignificant change in the
amount of slip where contact begins.
c. Both the carcass design and the type of tread pattern have a real ef
fect on the slip but the part played by the carcass design is of greater
importance.
308 MECHANICS OF PNEUMATIC TIRES

5.6 Tractive Forces on Dry Pavements


5.6.1 Introduction
The following list of five friction processes and a definition of terms is
offered in the interest of reducing the ambiguity surrounding the use of the
words "coefficient of friction."
a. The classical coefficient of friction, /i, is defined as being equal to FJ
W, where F3 is the force, tangent to the contact surface, applied to a solid
slider to initiate or maintain sliding and WK the force, normal to the con
tact surface, holding the sliding elements in contact. This latter is the
wheel load in the case of tires. Thus

b. Where no sliding occurs at any point, a force tangent to the contact


surface, F, may be applied to a perfectly rigid slider producing a uniform
tractive stress τ at the interface. We define F/W = θ, and obviously the
maximum values of F and θ are given by Fm« = Fs and #„„ = /t.
c. A force Fp, tangent to the contact surface, may be applied to a flex
ible slider such as a non-rotating tire, producing a nonuniform distribution
of tractive stress τ at the interface. This may occur where the nominal con
tact stress is nonuniform and/or where elastic constraints on the interfaces
are nonuniform. Using the findings of Mason, we. can state that a very
small value of tangential force will produce slip over at least a small part
of the interface. Fp thus always produces partial slip. As the value of Fp in
creases, a greater fraction of the interface slips. We define Fp/ W = <£ and
obviously Fp and <j>p at full slip equals Fs and ju, respectively.
d. A force Fr is required to roll a loaded tire on a level surface. When
control forces are not applied, this mode of operation is referred to as free
rolling. Fr is usually considered to be due to all of the hysteresis losses in a
deflecting tire, which results in moving the center of pressure toward the
leading edge of the contact patch. However, from the work of Reynolds
and others [67] it is apparent that even in free rolling of a cylinder tractive
stresses and sometimes slip may be induced in the contact region. In addi
tion, for tires, tractive stresses are developed in the contact patch because
distortion is required to change the curved surface of the tire to conform to
the flat road surface. For any complete model of the tire-road interface
friction, these tractive stresses must be superimposed upon the tractive
stresses caused by forces other than Fr.
e. A control force Fn either braking, traction, or lateral, may be applied
to a rotating tire. In this case, the tractive stress and slip distributions in
the contact region are more complex than described in paragraph (c)
above. This case is the subject of a number of papers [69, 70]. The point to
be made here is that (FJ^/Wis often referred to as the coefficient of fric
tion, confusing it with /*.
During braking, as the rotating wheel is slowed toward complete slip, an
increasing fraction of the contact patch slips. In this case, Fc approaches
CONTACT BETWEEN THE TIRE AND ROADWAY 309

F.„ and this is a result of increasing relative speed between the tire carcass
and road surface. On the other hand, Fp approaches Fs as a result of the
distance of movement of the center of mass of the slider, a very different
mechanism.
It is apparent that the operation of a tire at high slip will be directly in
fluenced by the factional properties of the tire rubber. It has been found
that [69] various factional conditions can influence tire elastic behavior at
small values of slip as well, where carcass deflections usually control tire
behavior. On the other hand, tire carcass properties strongly influence the
factional behavior of tread elements, particularly by way of controlling
tire-road contact pressure and local slip velocities. Because of these inter
actions, it is imperative that tire performance studies include detailed con
siderations of friction as well as of tire deformation and normal pressure
distribution.
One common method of acknowledging the combination of the several
events in the contact patch is to describe tire friction in terms of three
functional categories. They are dry friction, wet friction, and hydroplaning
[8, 70, 72]. The role of the road surface is often included under the descrip
tion, smooth, polished, abrasive, rough, well-drained, etc. The tire may be
characterized as being with or without tread and in some instances. the
tread material is given. Unfortunately, none of these descriptions is ade
quate for a factional characterization of a particular tire on a particular
road in a particular environment. A more fundamental approach follows.

5.6.2 Rubber Friction


For rubbing systems in general, there are several factors that contribute
to sliding resistance. The friction of tire rubber on practical surfaces can
be divided into at least four components, or causes. The naming of these
four components is arbitrary to some extent, and they are here separated
in terms of friction force, F, rather than the coefficient of friction /t:
«<Khesive) " M •

and each component is defined below. Each component is distinctive as


shown by simple experiments. Each component varies in a different way,
and affects tire performance in a different way, with variation in contact
stress, sliding speed, temperature, tire material, road surface profile, etc.
The several components of faction will first be described individually, for
clarity. In later paragraphs, the likely balance of the components will be
discussed in relation to a real tire.
The existence of the adhesive, or dry, component of rubber friction is
verified, and values are measured, by sliding a rubber specimen on a care
fully cleaned, smooth surface, such as glass. Such experiments show that a
thin smooth film of rubber about 100 A4 thick is deposited and remains at
tached to the mating surface. The sliding retardation force varies consid
erably with contact pressure, sliding speed, and temperature and is consis
tent with the view that the value of F,., is dependent on the viscoelastic
mechanical properties of the polymer [73], [72].
When a sliding body leaves a thin film or track, a friction test may be
merely a shear test. If this is the case, then /•'„,, depends on bulk mechani

4 Å-Anptrom units, 10 8 cm.


310 MECHANICS OF PNEUMATIC TIRES

cal properties of the polymer and not on adhesion kinetics. Other authors
[74-76] are of the opinion that adhesion kinetics is responsible for either a
part of, or the entire, adhesive frictional behavior. In their view, the rate of
making and breaking bonds controls the magnitude of Fad. Apparently
they assume that the bonds that form also break at the original interface.
One observation at least may be explained by surface kinetics, and that
is the need to slide rubber a few inches before a steady state value of fric
tion is measured. On the other hand, this may also be due to the existence
of surface films.
Whether Fad is controlled by adhesion kinetics or bulk mechanical prop
erties may in the end be a moot point. There is a possibility that these two
properties may be derived from the same source.
In any case, there is general agreement that /-',,d varies with sliding speed
and temperature as shown in figure 5.81. The plotted master curve (a) is
from reference [73] for an acrylonitrile-butadiene tire material with a t
(glass transition temperature) of -20° C (-4° F). The curves are shifted
using the WLF transform to show friction at various temperatures.
The location of the peak of the friction curves can be predicted from the
glass transition temperature of the rubber. For SBR rubber with tg ≃ -45°
C (-49° F), for example, the friction maxima at the temperature desig
nated in figure 5.81 would be shifted nearly two orders of 10 to the right
on the velocity scale. In the same manner, natural rubber with tg = -60° C
(-76° F) would have the peaks of the curves shifted nearly three orders of
10 to the left.
The main point to be gotten from figure 5.81 is that it appears that in

25
-20"C(-4eF) 20"C (68°F) 80'(I760F)

2.0

1.5

l.O

.5

-8 -6 -4 -2
Loo Sliding Speed ICm/secl
I ' I I
Sliding Speed IMiles/hrl 0.00022 0022 2.2

FIGURE 5.81 Coefficient of sliding friction plotted against speed in cm. Isec. and miles per
hour,
Curve (•) is the master curve for acrylomtnle-buladiene (unfilled) at 20* C (68" F). Positions or curve (a) shifted to —20*
C (-4° F) and 80° C ( 176° F) are shown. Curve (b) is for acrylonilrik-butadiene containing 50 parts of high abrasion fur
nace carbon panicles [73].
CONTACT BETWEEN THE TIRE AND ROADWAY 311

practice friction most often decreases as sliding speed increases. However,


tire-road slip causes surface heating which was not accounted for in figure
5.81. Heating will modify the effect of sliding speed, by an amount un
known at this time.
These curves "imply" that friction should increase as the ambient tem
perature increases. In practice, this is not found to the extent shown in fig
ure 5.81 and the cause of this discrepancy is not known. Perhaps the rea
son lies in the fact that other mechanisms besides the adhesive component
of friction are important in most real conditions. It should be noted that
practical values of /t are not as high as shown in figure 5.81. The reason for
this is also not known.
The values plotted in figure 5.81 are for kinetic or dynamic coefficient of
friction μk. It is tempting to extrapolate the curves to zero velocity to find
the static value, /*,. It would appear that /<, is smaller than /tt, which is op
posite to general expectations. The problem most likely lies in identifying
μs with absolute zero velocity, which may be reasonable for some sliding
systems involving more rigid materials. However, it is easy to see that in
order to achieve a practical sliding speed from zero velocity one passes
through low velocities and possibly through the velocity range where high
values of fi are measured. Thus the value of μ which must be supplied to
start sliding can be higher than the value of ju to maintain sliding. The lit
erature is not always clear on where the transition or peak value occurs on
the velocity scale. Reference [77] reports a few test results showing a
nearly linear decrease in ju with increasing sliding speed, with emphasis on
common automobile speeds. Hurry and Prock [78] report an increase in /i
with increasing sliding speed, but only up to speeds of 10 ft/min. (≃0.1
mph). There is no conflict between these two sets of data.
Friction induced vibrations are often caused by materials exhibiting re
duced friction with an increase in speed. Such vibrations are often attrib
uted to the "stick-slip phenonenon," which occurs in cases where μs is
greater than /t*. These vibrations are initiated by random disturbances ei
ther at sliding interface or in the machinery driving the friction experi
ment, and the frequency is determined by the dynamics of the system.
Standing vibrations are never found to occur at sliding speeds less than
that at peak values of μ (fig. 5.81). Vibrations occur to the right of the peak
and the vibrations are the more severe where the slope of the curve is
steepest. Careful analysis of vibration data shows that in many cases the
vibrations produce an oscillation in sliding speed between two finite val
ues, and both values are to the right of the peak of the curve. This cannot
be called the "stick-slip phenomenon"!
The entire explanation for friction induced vibration is not to be found
in the adhesion component of friction. More of the topic will be found in
the later discussion on the viscous component of friction.
The adhesion component of friction has been shown repeatedly [67] to
result from strong interface bonding at the small local sites of contact be
tween two surfaces. If bond strength is the same wherever bonds exist, the
force that resists sliding is proportional to the total of all of the minute
areas of contact. A number of factors control this true area of contact. We
could expect that the true area of contact of two atomic-ally smooth sur
faces would be equal to the nominal area of contact and fi would be very
high. However, for the usual surface which is very rough on the atomic
scale, contact would be limited to the highest proturbances on the sur
faces. The resulting true area of contact will depend upon details of the
surface profile, the magnitude of the average contact pressure and certain
312 MECHANICS OF PNEUMATIC TIRES

properties of the materials Estimates of true contact area are available for
two classes of material behavior. For materials that yield plastically the to
tal true area of contact A is

(5.14)
3Y

where W is the load per nominal unit area, Y is the yield strength of the
material, and A', is a factor of proportionality. For elastomers on rough
surfaces represented by an array of hemispheres we obtain the area of
contact from the Hertz contact stress equations:

A=K2^- (5.15)

where r is the radius of the hemispheres, E is the modulus of elasticity of


the rubber, and K2 is a factor of proportionality. The values of K1 and K2
vary with a number of conditions including surface roughness. For ex
ample, if surface roughness is represented by an array of spheres, an eight
fold decrease in the population density of spheres will decrease A by one-
half.
From eq. 5. 15 above it can be seen that the adhesion /t is proportional to
(average contact pressure)* on dry surfaces. This is observed in rubber and
other elastic materials as shown in figure 5.82. More general conclusions
on this subject can be reached after discussion of the remaining com
ponents of rubber friction.
/•",,,., is a retardation force available when sliding a rubber specimen on
perfectly lubricated surfaces with smooth bumps or proturberances [67].
The retardation force is due to the partial irreversibility of deformation,
i.e., damping loss of the rubber caused by passage of the bumps or protu
berances. This component of friction is not significant until there is slid
ing. The need to attribute FM to the damping loss of tread rubber is not
often disputed. Attempts have been made to connect dry f rictional behav-

3 2.1
1
o
ac.
PO b X
u.
L. — u>
O
H
Z —b
Ul

ili 0.5
o
o
10 15 20
NORMAL APPARENT PRESSURE (KG/CM')

FIGURE 5.82 Coefficient offriction vs. normal apparent pressure [17].


CONTACT BETWEEN THE TIRE AND ROADWAY 313

-3-2-101234567
LOG FREQUENCY OF VIBRATION

FIGURE 5.83 Typical curvefor damping loss as a function of log frequency of vibration from
a vibrating reed test.

ior with the damping properties of rubber. More work is required to re


solve this point.
Damping loss varies with frequency of deformation as shown schemati
cally in figure 5.83. It is reasonable to expect that sliding over a regular ar
ray of bumps produces a vibration in the rubber, the frequency of which
would be related to sliding speed. Where the bumps are 1/4 inch (0.6 cm.)
diameter spheres and the rubber is acrylonitrile-butadiene with tg at
-20°C (-4°F), the retardation force due to damping alone varies with
sliding speed as shown in figure 5.84.
If topographical features of a real road are included, it is possible to
synthesize a more general deformation component loss curve. The large
texture of a road surface is approximated by the 1/4 inch (0.6 cm.) diame
ter spheres just mentioned [80]. A second finer level of texture exists which
has a radius of about two orders of 10 less than the course texture. Finally,

uj tr
=> uj
Q 00

s
is
Q Q

J I
-3-2-10123456
LOG SLIDING SPEED (M.PH)

FlOURE 5.84 Sliding retardation force due to damping loss in rubber when sliding over an
array of spheres 1/4 in. (0.6 cm.) diameter.
314 MECHANICS OF PNEUMATIC TIRES

/Very Fine Texture


/Fine Texture
OI6W Coarse Texture

J I J I
-3-2-10 I 234 5 6
LOG SLIDING SPEED

FIGURE 5.85 Schematic representation of the sliding retardation force due to damping loss in
a high loss rubber when sliding over a three order array of spheres, b — 1/4 in. (0.6 cm.), b X
1<T2, b X 10-3'.
The three texture* are not known to produce equal effects.

a third very fine level of texture exists at about three to four orders of 10
less than the course texture. The friction performance of the three textures
is shown in figure 5.85 together with the approximate theoretical maxi
mum value of Fdef ≃ 0.16 for a very high loss rubber.
If in the place of discrete steps in texture size we assume a continuous
distribution, a very broad curve could be plotted as shown by curve a in
figure 5.86. On the other hand, if one range of texture size is missing from
the road surface, the retardation force curve could be altered as shown by
curve b in figure 5.86. This curve represents an alteration of the fine tex
ture of figure 5.85 as occurs on polished roads. This effect is more fully de
scribed below.
Curve b of figure 5.86 may also serve as a basis for explaining the de
crease in wet friction as temperature increases. The curves in figure 5.86
shift to the right with an increase in temperature, which has the effect of

OI6W

I i ; i I I j i
-4-3-2-10123456
LOG SLIDING SPEED

FIGURE 5.86 Sliding retardation force due to damping loss in a high loss rubber sliding on
"a", a surface of a continuous distribution ie., continuous effectiveness of sphere sizes within the
range offigure 5.84 and "b" with size 6 X 10~2 missing.
CONTACT BETWEEN THE TIRE AND ROADWAY 315

lowering Fder at moderate to slow sliding speeds. The effect is found by


measurement to be the greatest on well-polished roads, which also fits the
curves of figure 5.86.
An interesting finding is that the maximum friction coefficient available
from the deformation component of friction depends upon the population
density of the fine protuberances on the road surface. For example, for
most tire rubber, with an average tire-road contact pressure of 30 psi, FM
≃ 0.07 W when sliding over a tight-packed array of spheres, while F^, ≃
0.09 W when sliding over the same array with alternate spheres removed.
The theoretical maximum F^t ≃ 0.16 W, as previously indicated.
The deformation component of friction is difficult to separate from
other components on complex surfaces. However, in spite of the low theo
retical maximum value of ju = 0.16, this could be the primary friction
mechanism at high sliding speeds on films of water. It would be expected
that F,,ef would be primarily influenced by the damping loss (rebound)
properties of the rubber. This has been found to be true for tires on wet
roads although the effect of rubber hardness confused the picture to some
extent [82].
Increased rubber hardness would have the effect of increasing contact
pressure and when contact pressure is increased by increasing the load,
FM increased [80]. On the other hand, harder rubber often has lower
damping loss (higher rebound) which could more than offset the effect of
contact pressure.
The viscous component of friction force Fvis is defined to take account of
the existence of a layer of either adsorbed or liquid species between the
tire and the road surfaces. It is presumed that this layer is thick enough to
significantly reduce direct bonding, or adhesion, of the tire rubber to the
road material. The uniqueness of the viscous component can be demon
strated by rubbing a rubber slider on a glass that has been carefully
cleaned and exposed to water vapor. A high friction force may be mea
sured on this glass surface, with F approaching W/2. In this experiment,
there is probably no Fdef to confuse the picture because of the smoothness
of glass, and no thin film of rubber remains attached to the glass.
It would appear that the experiment described above, producing F ≃
W/2, is one extreme example of a fluid film. However, this may, in fact, be
the most common condition of tire-road contract, due to the adsorption of
water vapor on tire and road surfaces.
The thickness of these adsorbed water films is not known. The closest
approximation can be made from the work on water adsorption on lime-
soda glass, a material of the same class as some road stones [82]. A surface
water film, apparently of 105Å (0.4 × 10-6 in.) thick, may persist in a dry
atmosphere at 23° C (41° F), and 55 Å thick at 215° C (390° F). Very
likely much thicker films would exist at lower temperatures and higher rel
ative humidities. In addition, when two mating surfaces each have an ad
sorbed film, the total separating film is still thicker. More careful work is
needed in this area.
Another important unknown quantity is the description of the appropri
ate properties of the film separating the rubbing surfaces. A thin film of
lubricant may behave as a fluid of greater apparent viscosity than the bulk
viscosity [83]. This effect is thought to be due to the nature of the bond be
tween a liquid and solid, which is manifested by the tenacity with which
liquids remain adsorbed to a solid. There is some criticism of this view,
but it appears to be a criticism of form rather than substance.
316 MECHANICS OF PNEUMATIC TIRES

Other recent works shows an opposite effect [85]. Electrical double lay
ers in an electrolytic liquid film between rubber and glass were observed
to maintain a larger equilibrium separation between the sliders than can
' be accounted for by viscous or inertia effects. Thus, a value of /"vis was
measured for a fluid film thicker than normal for the contact stress, pro
ducing a very low value of /i = 0.04 without hydrodynamic lift.
In principle, at least, FvU can be sliding speed dependant in such a way
as to encourage factional vibrations. The proper conditions are achieved
by rounding the edges of rubber elements so that hydrodynamic lift in
creases with velocity. These conditions are apparently met when a tire
slides on a wet polished concrete surface. A high frequency vibration
(squeal) can be heard, particularly at speeds below 15 mph, and it origi
nates in the tread region of the tire.
FU^ is a component of friction which takes account of the observation
that some solid surfaces tear particles from the rubber. These particles
usually do not remain attached to the mating surface, and this mechanism
can occur on contaminated surfaces where values of /i may be low. In
some cases, where values of /i are high, it is easy to explain the tearing of
particles from a rubber surface by high traction stress, in combination
with contact stresses, which cause fracture in the rubber. Likewise, it could
be argued that /•',„, is already accounted for under one of the other three
components of friction. On the other hand, the deformation and thick film
viscous components of friction would not ordinarily be expected to pro
duce large wear fragments.

5.63 Tire Traction under Dry Conditions


The operation of tires on roads involves some combination of the sev
eral mechanical classes and components of friction. Research has not yet
reached the stage where it is possible to mathematically express either the
distribution of slip velocities in the contact patch, or the traction stress dis
tribution due to the components of friction on tire and road surfaces of
undefined surface geometry. It will therefore be necessary to qualitatively
describe some of the events in the contact patch under various operating
conditions, and to further define some terms. In these descriptions, very
many details are omitted, in the interest of maintaining the broader and
more unifying view.
Dry Friction
Dry friction is one phenomenological category of tire operation that is
thought to be simple, and it usually produces high friction. In fact, it is a
wide range of conditions. A common but probably inaccurate assumption
is that dry friction involves primarily the adhesive component of friction.
Several interesting coincidences seem to bear this out, but the situation is
far from resolved. For example, begin with curve a of figure 5.81 for dry
friction of an unfilled rubber where a film of rubber was not left on the
glass surface. Reference [73] shows the peak value of friction for the same
rubber at nearly the same sliding speed as that of reference [72] in an ex
periment that produced a thin film of rubber attached to the glass. Refer
ence [72] also reports data for carbon filled tire rubber, which is a curve
that overlaps the curves of figure 5.81 except that the peak of the curve is
CONTACT BETWEEN THE TIRE AND ROADWAY 317

replaced by a plateau showing a maximum sliding coefficient of friction


near 1.2, a value that has been reported for tires on dry roads. On the
other hand, a μ of 1 .2 is not proof of the adhesion component of friction
alone. Reference [73] measures a maximum /i of 3.2, but reference [85]
shows a value in excess of 10. It is likely that these differences are due to
complex surface films.
A "dry" surface may be covered with adsorbed water up to the limit of
invisibility at about 500 Å (2 × 10~* in.) or Vi micron thick. A reduction in
braking friction is possible on surfaces exposed to cool moist air without a
visible water film [86]. This could be true with a 500 Å thick film.
It can be estimated, using the simple Newtonian viscosity equation for
drag force per unit wetted area, f = TJP//JQ, that the viscous drag of a water
film of about 10-6 cm. thick between two flat surfaces may produce Fvis in
excess of 0.25 W at a sliding speed in the order of 1 mph. For thinner films
or higher velocities, Fvu would be higher. This line of reasoning ends with
the comment that (Fc)max for tires does not often exceed W, in which case ji
= 1 . Conversely, it has not been proven that some penetration of the water
film [83] does not occur on very sharp asperities, thus producing some ad
hesion friction. However, it is not necessary to invoke adhesive friction to
account for the high values of tire traction measured on so-called dry sur
faces.
At larger slip velocities, such as during braking or severe cornering, the
thin water film would probably be heated and possibly boiled away to
produce more adhesive friction. A simple calculation shows that where F
≃ 0.5 W, average contact pressure is 30 psi, and due account is taken of the
heat of adsorption of water to glass, adiabatic heating, and boiling of a
water film 10 '' cm. (0.4 × 10 6 in.) thick takes place by rubbing a distance
of from one to two inches. Thus, for a typical automobile tire a slip of
about 15 percent would boil such a water film away. For larger slip ratios
or for thinner water films, the water would be removed in the leading part
of the slip region in the contact patch, producing dry or adhesive friction
over the remainder of the contact patch. Traction stresses could, therefore,
decrease as slip velocities increase (see fig. 5.81) the opposite of what
would be expected if a viscous film were to persist. Under still more severe
conditions, the surface temperature of the rubber could reach the soft
ening or reversion temperature where rubber is easily removed from the
tire surface [53].
Rubber Properties
There is no tire rubber that produces high friction on all surfaces, under
all conditions. The effect of base polymer and of compounding ingredients
on frictional properties appears to depend upon a number of other condi
tions describing the sliding process, including the nature of the opposing
surface, the lubrication, and the load. For example, increasing carbon
black content has been reported to increase friction coefficient on an abra
sive surface but to decrease it on a smooth surface at low loads.
Soft compounds usually give higher coefficients of friction than do
harder compounds on smooth surfaces under low loads. In line with this
observation the effect of rubber plasticizer is to produce softer rubber,
while increased state of cure produces harder rubber and lower friction.
The effect of rubber hardness on its frictional properties is complicated
by a reduction in hardness of the rubber layer in immediate contact with
the opposing surface as a result of severe mechanical flexure during sliding
[92]. The hardness of a layer of filled rubber in immediate contact with an
318 MECHANICS OF PNEUMATIC TIRES

abrasive surface is very close to that of an unfilled rubber. For tires, any
effect of hardness increase, caused by addition of fillers, on frictional force
must therefore be attributed to the effect of the properties of the rubber
below the surface layer. It is well known that these substrate or bulk prop
erties can be important, since in comparing the friction of thick and thin
rubber membranes on ice the thin specimens exhibited consistently lower
coefficients of friction than did the thicker ones.
The influence of rubber properties is complex. However, in summary, it
can be said that the choice of base polymer and of the ingredients used in
compounding can affect frictional properties not only through their influ
ence on such physical characteristics as hardness, damping loss [81] and
surface roughness, but also by changes in chemical adhesion and the ease
of surface contaminations through bleeding of pigments or adherence of
extraneous contaminants. Furthermore, the compounding necessary to
achieve maximum friction may need to be specific to the opposing surface.
In view of the several constraints on the practical range of tire rubber
properties the available range of friction is not large. By far the most im
portant variables are the nature of the road surface and the tread pattern.
Longitudinal and Lateral Tire Slip and Slide
The frictional forces between a tire and road vary with the amount of
deviation from straight line free rolling operation. For the classical rigid
wheel rolling on a flat plane any slip whatsoever is complete sliding. How
ever, the flexible tire structure can deviate from straight line free rolling
without complete sliding at the tire-road interface. Although the exact me
chanics of this behavior have not been worked out the results are well
known. Some are summarized below.
The behavior of the tire in braking is shown in figure 5.87. The data are
presented in terms of brake torque coefficient and percent slip. The brake
torque coefficient is a value which is, to within a few percent, the ratio of
braking force to normal load. Percent slip is defined as
Wr-W,
(100) = percent slip
Wr

20MPH-I.35
40MPH- 1.50
60MPH-I.45

20 40 60 80 100
PERCENT SLIP
FIGURE 5.87a. Brake torque coefficient vs. slip-wet asphalt (94J.
CONTACT BETWEEN THE TIRE AND ROADWAY 319

1.2

20 MPH -1.30
40 MPH -1.20
60 MPH -1. 10
t-- 1-
ffi

20 4O 60 80 100
PERCENT SLIP

FIGURE 5.87b. Brake torque coefficient vs. slip-dry asphalt [93].

where W, = rotational speed of a free rolling reference wheel and Wt =


rotational speed of the test wheel.
The behavior of the tire in side slip can be seen in figure 5.88 giving
data for a number of tires, road surfaces, and vehicle speeds [94]. Data for
combined braking and side slip are found in the same paper. However, a
clearer picture of combined effects are found hi figure 5.89.
Several attempts have been made to mathematically express the forces
available in all possible directions and degrees of slip. This usually takes
the form of a "friction circle" or "friction ellipse." In the case of an as
sumed friction circle, one hopes that it would be possible to predict total
behavior from a simple braking test. Unfortunately, available data [31] do
not show this to be practical. An expression of the form R2 = B2 + S2 is
often used, where R is the resultant horizontal force, B is the braking
force, and 5 is the side force. There is not a detailed discussion of this

(c)

40 MPH -20MPH

PEAK/SLIDE RAT 10
20 MPH -1. 50
40MPH-I.40
60 MPH -1.70

20 40 60 80 IOO
PERCENT SLIP
FIGURE 5.87c. Brake torque coefficient vs. slip-wet portland concrete [94J.
320 MECHANICS OF PNEUMATIC TIRES

20 MPH -1. 25
40MPH-I.I5
60MPH-I.05

0 2O 40 60 80 IOO
PERCENT SLIP

FIGURE 5.87d. Brake torque coefficient vs. slip-dry portland concrete [93].

topic in hand. It would he surprising if a single simple expression of the


above form could be applied to all cases. For example, the effect of speed
on the above expression is known to depend on tread design. Further
more, it makes some difference whether the data used are peak values of
braking force and side force, or sliding values. Speculation on this subject
may be near an end since the advent of high quality test devices [95] to
measure force values at all combinations of braking and side slip.

5.7 Tractive Forces on Wet Pavements


5.7.1 Introduction
Any difference that exists between the traction values measured under
wet and dry conditions for a given tire-pavement combination is due to
1.2 I
PEAK/SLIDE RATIO
20 MPH DRY -1.55
40MPH DRY -1.30
20 MPH WET-1.30

SLIDE

,20 MPH DRY ,40 MPH DRY

0 20 40 60 80 100
PERCENT SLIP

FIGURE 5.87e. Brake torque coefficient vs. slip-ice [93].


CONTACT BETWEEN THE TIRE AND ROADWAY 321

10* 15" 20" 1C1 15* 20'


SLIP ANGLE OC SLIP ANGLES

FIGURE 5.88. Effect of slip angle on sideway force coefficient J94].

Longitudinal
Friction
Coefficient

20 40 60 80 100
WHEEL SLIP PERCENT

FIGURE 5.89. Friction coefficient vs. wheel slip [95].


322 MECHANICS OF PNEUMATIC TIRES

the persistance of a fluid film in portions of the footprint area. The pres
ence of a fluid film under wet conditions reduces the dry contact area be
tween tire and pavement resulting in lower traction values since only in
the areas of dry contact can a significant level of traction be developed.
Actually, even under dry conditions, there is no such thing as a perfectly
dry contact region between tire and pavement; fine water films con
taminate all surfaces when the humidity is nonzero. Therefore, when ref
erence is made to a dry contact area, this implies a region in which the
fluid film thickness has been reduced to that thickness present on what is
casually and technically incorrectly termed dry pavement.
Under wet conditions the contact patch can be considered to be divided
into several different regions based on the state of the fluid film [8]: first, a
region of unbroken water film, second a region in which partial break
down of the fluid film has occurred with the tread rubber beginning to
drape over and establish contact with pavement asperities, and third a re
gion of essentially dry contact in which the fluid film has been nearly ex
pelled from most tread-asperity interfaces. The major portion of traction is
developed in this latter region. As previously discussed both the size and
shape of these footprint regions depend on a large number of factors. This
three zone model is valid insofar as it delineates regions in the contact
patch where the traction stresses are controlled by different mechanisms.
The third zone cannot be taken as absolutely dry however, since this im
plies a very large normal stress to reduce the fluid film to zero, and is un
necessary in any case. High traction can be achieved in very thin water
films.
In general, factors affecting the wet traction performance of a tire pave
ment combination can be grouped under four major headings: the tire, the
pavement, the fluid, and operating conditions [97], [98]. Albert and
Walker [8], by summarizing the results of many studies, were able to make
general statements, shown in Table 5.5 of the level of variability in wet
traction performance due to several of these factors.
Beside the several excellent general discussions of factors affecting wet
traction, [98, 99, 100, 101] numerous reports have been published detailing
the effects of individual parameters. References are given in Table 5.6 for
a number of representative papers.

5.7.2 Thick Water Film Effects


Thick film wet traction exists when the major groove network in the
tread pattern becomes flooded, so the flow hydrodynamics in the groove
network in the contact area becomes the dominant factor in traction. Such
a situation commonly occurs on puddled and flooded pavements, where
there is almost always a reduction in the effective contact area between the
tread surface and the roadway as compared to dry pavement levels. This is
because the mass and viscosity of the fluid cause it to resist being dis
placed from between the tire and the pavement, thus generating lift forces
on the tire surface as it moves through the fluid film. When these lift forces
under thick film conditions become great enough to completely support
the load on the tire, contact with the road is no longer made and dynamic
hydroplaning is said to occur. Since the only contact with the road is
through the fluid film, almost no braking or steering forces are available.
This is well illustrated by data obtained by Home and Joyner [33] shown
in Figure 5.90.
CONTACT BETWEEN THE TIRE AND ROADWAY 323

TABLE 5.5 Factors influencing effective brakingfriction between tire and wet road (100 mph
maximum |8|
Level of
variability
Factors due to factor
considered
Tire factors
1. Tread pattern design. Scale of effect recorded during assessment Up to 4:1
of practical tire pattern designs.
2. Tread materials. Maximum effect due to tread material changes Up to 1.5:1
within range of materials currently available

3. Patterned tire. Smooth tire Up to 8:1


Road factors
1. Road surface characteristics. Scale of effect recorded during Upto5:l
assessments of practical tire designs
2. Water depth. Effect of variation in depth of water film covering Upto3:l
a road surface in the range 0.05-0.20 in.
Vehicle factors
1. Speed. Reduction in braking force coefficient due to an increase Up to 10:1
in speed from 30 to 80 mph (patterned tires only)
2. Braking system. Perfect nonlocking system versus locked wheel Up to 3 : 1
braking •
Note: Cases of braking force coefficients below a value of 0. 1 have been excluded.

TABLE 5.6 Factors affecting wet traction


Variable Factor Reference
Tire Carcass Construction
Type, size, aspect ratio 114, 176, 177, 179
Tread Pattern Design 112, 113, 136, 178-180
Tread Depth 112,115,123,181-183
Tread Compound 142, 184
Inflation Pressure 112, 116, 123
Pavement Surface Texture 118,147
Micro-texture
Porosity 119
Aggregate Wear Characteristics 120, 121
Fluid Density 33,113,136
Viscosity 33,113,136
Film Depth 116,123,124,139,140
Operating Velocity 139, 140, 178
Load 112,123,179
Percentage Tire Slip 185
Temperature 123, 179, 186
Traffic Density 187, 188
Site Design 189, 190
324 MECHANICS OF PNEUMATIC TIRES

Jextured concrete
looth concrete
, Predicted
_ K hydroploning
V tr-ioari t»_

20 40 60 80 ICO 120 0 20 40 60 80 100 120 140


GROUND SPEED, VG.mph
(b) Smooth tread (b) Smooth tread
Water depth =0.04 in. Water depth = 0.3 in.

FIGURE 5.90. Effect of tread design, runway surface texture, and speed on the locked wheel
friction coefficient, \>.,kid, developed on a flooded runway.
6 50 - 13 puienger car lire, vertical load - 835 Ib. tire picture - 27 Ib/m-' |3)|.

As might be expected, the role of the various design variables is some


what different in thick film wet traction than in dynamic hydroplaning,
since in the pure hydroplaning case the tread surface cannot contact the
pavement so that tread compound and pavement micro-texture have little
influence, while they have considerable effect on the wet traction forces
available prior to hydroplaning.
Full dynamic hydroplaning occurs when the amount of fluid encoun
tered by the tire exceeds the combined drainage capacity of the tread pat
tern and the pavement macrotexture. It occurs in deep fluid layers, usually
greater than 2.0 mm, where fluid mass effects are dominant. As a tire
moves through a thick fluid layer, the momentum of the fluid is changed
which causes a reaction force on the tire tread surface. The pressures so
built up on the tire surface are sufficient to bend the tread region inwards
and upwards from the pavement allowing a progressive persistance with
increasing speed of the fluid film from the front, through the middle along
the centerline, to the rear of the footprint. Since tire deformation is related
to the imbalance between the external and internal pressures acting on the
tire carcass, Home [33] discovered that the following formula gives the ap
proximate speed at which a tire will dynamically hydroplane.
VH = 1.8V?" (5.17)
where V,, is the threshold speed for hydroplaning in m/s, and p is tire in
flation pressure in kilopascals. Comparison of this equation with experi
ment is given in Figure 5.91 and is seen to be excellent. The approximate
speed levels at which various ground vehicles will be susceptible to dy
namic hydroplaning are shown in Figure 5.92. Note that eq. 5.17 applies
only when the requirements are met on the fluid depth, tire tread pattern,
and pavement macro-texture. In general the potential for dynamic hydro
planing is increased by little pavement macrotexture with few inter
connected drainage paths, shallow tread depth, poor tread pattern design,
CONTACT BETWEEN THE TIRE AND ROADWAY 325

>20O

V = 9fp / LOAD
PER

120
7• i TIRE
12X66 (VII)
TIRE,
LB
9,400
• 4M»I3.0(VII) 22,000
VEHICLE

Aircraft
Aircraft

^ 80 «d
X • 1 700-20(111) K3.000
(
•xiVvn)
C-123
Aircraft
660

Jrt- J
io" • Transport

40 o 250-16
9,100 Four-WnMl
•"'" 38VI2.3') Bogi«
570-15 929

10 20 40 60 100 200 400


TIRE PRESSURE, p,lb/in2

FIGURE 5.91. Experimental and calculated tire-hydroplaning velocities [33J.

deep fluid layers, wide, low inflation pressure tires, and high vehicle
speeds [6], [33], [106].
It is convenient to present the influence of the various variables on dy
namic hydroplaning speeds by use of a Table 5.7, which attempts to con
dense the results of a great deal of research into an extremely concise
form. In doing so, however, many of the nuances of the rather extensive
literature are lost, and the serious reader should avail himself of the origi
nal papers used here as sources. For example, the influence of tread depth
is quite strong in hydroplaning. As tread depth decreases the speed and
water depth necessary for hydroplaning are both reduced, as is shown in
Figure 5.93. The results of one investigation on the interaction of inflation
pressure, load and speed is given in Figure 5.94, which clearly demon
strates that the onset of hydroplaning can be load dependent as well. How
ever, the relationship between load and hydroplaning potential or wet
traction capability is quite complex [7, 106, 111,1 12]. Increases in load un
der dry conditions produce increases in the length of the contact region

20 30 40 60 80 100
TIRE INFLATION PRESSURE , p , Ib/ in?

FIGURE 5.92. Susceptibility of some ground vehicles to hydroplaning. [6],


326 MECHANICS OF PNEUMATIC TIRES

TABLE 5.7 Factors influencing dynamic hydroplaning


general references (7, 33, 96, 106, 111-113)

Comments References
Open Variables
Speed High speeds induce hydroplaning 33, 99
Inflation Higher inflation retards hydroplaning 33, 111, 116
Water Depth Approx. 2mm minimum required. 33, 1 16, 123
Above that little influence.
Load Higher loads increase hydroplaning speed 106
Tire Variables
Tread Depth Small tread depth promotes hydroplaning 106, 111, 112, IIS
Contact Patch Width Wider contact patch hydroplanes at lower 106, 111, 112
speeds
Construction
Radial Less tendency to hydroplane due to less 7, 96, 106, 1 14
groove closure, more effective tread
patterns
Bias More tendency to hydroplane for reasons
above
Tread Pattern
Wide straight Reduces hydroplaning tendency 7, 96,
grooves
Narrow Ribs Reduces hydroplaning tendency 106, 1 12
Tread Compound Little influence
Pavement Variables
Good Macrotexture Retards hydroplaning 117,118
Good Microtexture Little influence

and increases in the normal contact pressure under the outer ribs of the
tread. On wet surfaces, increased contact length helps to raise the velocity
at which total hydroplaning will occur. Increased normal contact pressure
under the outer ribs helps these ribs to more rapidly penetrate the fluid
film, which helps to raise the hydroplaning velocity. However, it also tends

Water film: O.I to 0.15"


Harsh textured asphalt surface
185-15 Automobile tires- Radial ply

Hydroplaning
0.25 0.50 0.75 1.00
GROOVE DEPTH WORN
RATIO
GROOVE DEPTH NEW

FIGURE 5.93. Effect oftread pattern groove depth on braking grip, "flooded" road conditions
IS].
CONTACT BETWEEN THE TIRE AND ROADWAY 327

0 10 20 30 40 50 60
IP INFLATION PRESSURE, psi
185-15 Automobile Tire-Rodiol Ply
Steel Drum Surface
0.05" Water Film
500 to 1500 Ib load

FIGURE 5.94. Effect of load and pressure on hydroplaning speed [106].

to close the tread pattern, especially in this region, which would act to
lower the speed at which hydroplaning would occur. Finally, the greater
variation in contact pressure between the middle and the outer ribs that is
produced by increasing the load will act to channel more water to the cen
ter of the footprint, which serves to promote hydroplaning. The net effect
that load will have on hydroplaning potential and wet traction is thus not
always clear. For instance Stocker and Lewis [111] found that increasing
load reduced the hydroplaning speed of a full tread depth tire and in
creased the hydroplaning speed of a smooth surfaced similar tire. The in
crease in groove closure accompanying an increase in load would thus ap
pear to have been the dominant effect in their experiments. Allbert [106]
found that a close relationship exists between the effects of load and infla
tion pressure on hydroplaning speed: the minimum speed for fully devel
oped hydroplaning can be estimated with eq. 5.17 only if tire load and in
flation pressure are adjusted so as to maintain a constant tire deflection.
Staughton and Williams [116] found that changing the load from 890 to
2265 N had little effect on the minimum speed for dynamic hydroplaning
of radial tires for inflation pressures ranging from 50 to 330 k Pa on
smooth concrete surfaces with a water depth of 9.5 mm. However, for both
full tread depth and bias tires worn smooth, they found, under the same
test conditions, that hydroplaning speed increased with increases in load,
these increases in speed being most significant for inflation pressure less
than 120 k Pa.
Site design and traffic density can have significant effects on the poten
tial for hydroplaning. Yeager [7], Gallaway, et. at. [126] and Dun lop, et.
al. [127] among others have related poor choices of roadway crown, cross
slope, grade, and radius of curvature to high pavement water depths and
thus to increased potential for hydroplaning. Increased traffic density re
duces the chance of hydroplaning through the path clearing action of the
preceding tires although at the same time it increases the hydroplaning po
tential due to the increased pavement rutting accompanying the greater
traffic volume, thus permitting puddling of water [128].
Highway and runway grooving has been found to be of great benefit in
reducing hydroplaning tendencies. Transverse grooving is widely used in
328 MECHANICS OF PNEUMATIC TIRES

^Impervious mix

Porous friction course concept

(b)

FIGURE 5.95. Groove and porous friction course concepts.

airport runways, mostly by sawing or flailing grooved patterns in existing


surfaces. Typical designs for such grooves are shown in Figure 5.95a. Lon
gitudinal grooves have also been adopted for many dangerous locations
on highway surfaces, and these have been successful in reducing accident
frequencies. In both cases the grooves act as drainage channels, which
carry water or trap it beneath the contacting surface of the roadway, thus
preventing build up of water films sufficient for hydroplaning.
More recently, there has appeared the concept of the porous friction
course, a layer of open graded asphalt ic agregate which is porous to rain
fall, and which allows the water to move vertically through it. This is illus
trated in Figure 5.95b. This design has also been successful in reducing or
eliminating hydroplaning by carrying excess water away from the surface,
as described in ref. [1 18, 1 19].
Because of the great danger associated with hydroplaning much re
search has been focused on developing means for predicting the circum
stances essential for its occurence. With such information appropriate ac
tion can then be taken to minimize the probability of the phenomenon.
Tire design factors related to hydroplaning have been previously dis
cussed. Once tires have been selected, a driver's only control over the
probability of dynamic hydroplaning is to keep his speed below that in
dicated by eq. 5.17 and to recognize that worn tires aggravate the phe
nomenon.
Dynamic hydroplaning may be thought of as one end of the spectrum
of speed and deep water film effects. It can easily occur in aircraft opera
tion but less easily in road vehicles due to their usually lower speeds. The
remainder of the spectrum of speed and water film effects is made up of a
region where at least a portion of the contact area of the tire has sunk
through the water film and makes effective contact with the asperties of
the road surface. Here, tire tread compound and road micro-roughness be
gin to play an important part since they determine the magnitude of brak
ing or cornering force transmitted through this zone, which is the only di
rect contact with the road. For this reason it would be expected that tire
operation in thick water films would have two basic objectives:
(a) The avoidance of hydroplaning by use of the principles previously
discussed, such as shown in Table 5.7.
(b) Maintenance of large tractive forces in the effective contact area be
tween tire and road surface.
For the most part the variables listed in Table 5.7 have the same effects
on thick film wet traction as they do upon the onset of hydroplaning, and
Table 5.7 may also be used in that way. However, much of the problem of
tire operation in thick water films lies in the transition region between full
CONTACT BETWEEN THE TIRE AND ROADWAY 329
0.8
Tire Pressure, kPo
330
» •» 215
• • 110
o o 55
Load: 1 335 N
Water Depth: 9.0 mm
Surface: Smooth Concrete -

20 40 60 80 100 120 140


SPEED , km/hr

FIGURE 5.96a. Effect of inflation pressure on locked wheel braking force coefficients
produced by a bias-ply automobile tire in deep water [116].

effective dry contact and complete hydroplaning. Within this region many
variables can combine to make one set of circumstances better or worse
than another, and in most cases test data must be used to actually deter
mine quantitative values for breaking and cornering. For example, in
creases in inflation pressure have been found to improve thick film wet
traction [33, 111,1 16]. This is seen in the case of a bias tire in Fig. 5.96a
and for a radial tire in Fig. 5.96b. This positive effect was found to hold
for a wide range of tire loads and pavement surfaces. That this effect is re
lated to the increased carcass stiffness in the footprint region can be de
duced by a comparison of the data in Fig. 5.97 for a worn bias ply tire
with that in the previously mentioned figures.
Tread pattern design and depth can have a major effect on thick film
wet traction performance. Data reported by Nordstrom [124] and Allbert
and Walker [8], among others, indicates that elements of tread design es
sential for good thick film performance are wide, straight longitudinal and
lateral grooves for channeling the bulk of the fluid through and out the
sides of the footprint and narrow ribs and small block elements to reduce
the distances water must flow to enter the grooves, these being the same as

Tire Pressure, kN/m*


330
4 215
—• 110
—o 55
Lood: 1335 N
Woter Depth; 9.0mm
Surfoce: Smooth Concrete

20 40 60 80 100 120 I4O


SPEED, km/hr

FIGURE 5.96b. Effect of tire pressure on the locked wheel braking force coefficient (fully
patterned radial) [116).
330 MECHANICS OF PNEUMATIC TIRES

CBRAKING
FORCE
OEF ICIENT _P
*
aM>
o '

Tire Pressur e,kN/m2


330
pi c
£
P
1 • 1 10
•> a 55
Loa d 1335 4
er depth 9.0mm
N Sur face snx oth cone rete

^\
..J-

) 20 40
^
60 80
:Z.
foO
—•
120 14
SPEED, km /hr

FIGURE 5.97. Effect of tire pressure on the locked wheel braking force coefficient (crossply
worn smooth) [116].

for an effective anti-dynamic hydroplaning tread design. This is shown in


Fig. 5.98. Open tread pattern designs with reduced real area of contact un
fortunately may show reduced traction under dry and damp pavement
conditions, as well as being noisier and wearing more quickly. A minor
negative effect is the increased vertical flexibility of the tread region.
Staughton and Williams [116] demonstrated this indirectly by showing
that increases in inflation pressure would have little effect on the thick film
wet traction values of unpatterned bias tires of full tread depth, since these
already had reduced vertical flexibility in the tread region. See Fig. 5.99.
The principal effect of tread compound on thick film wet traction values is
its major role in determining the local friction coefficients and thus the
traction generated in those portions of the footprint region where "dry"
contact has been established. In general, increasing the hysteresis and re
ducing the hardness of the tread compound will raise wet traction values.
Table 5.8, which shows the effects of rebound (hysteresis) and hardness on
wet road stopping distance, is taken from the work of Peterson, Eckert,
and Carr [142] who have made an extensive review of the effects of tread
GROOVE
RIBS GROOVES WIDTH
(No.) (No.) (In.)
9 4 0.2
K>
t>
^O
i*
con 6 5 0.2 1
7 6 0.2
FORCE
BRAKE
COEF ICIENT
P
p—
^

'~~~7Rib] Peak
•-^-- < ^ ^e Rib ;°r9
^"^SRib j vo1"**
~-^7Rib I Locked
^^6Rib > wheel
,\5Rib j values
) 10 20 30 40 50 60 70
SPEED, miles /hr

FIGURE 5.98. Results of braking tests on tires with four, five and six equal width
circumferential grooves ffj}.
CONTACT BETWEEN THE TIRE AND ROADWAY 331

Tire Pressure, kn/m2


330
215
HO
55
Lood 1335N
Water Depth: 9.0 mm
Surface: Smooth
Concrete

40 60 80 100 120 MO
SPEED, km/hr

FIGURE 5.99. Effect of lire pressure on the locked wheel braking force coefficient 1116] (full
tread crossply smooth).

compound on tire traction. This subject is treated at length in section 5.6


on dry traction. The greater the tread depth, the greater can be the values
of water depth and vehicle speed and the fewer can be the drainage chan
nels between elements of pavement macro-texture, without the drainage
capacity of the tread groove network being overtaxed and dynamic hydro
planing occurring. Gallaway, et. al [126] performed an extensive study in
which vehicle speed, tire type, tire pressure, surface texture, water depth,
and tire tread depth were varied. He found that in all thick film cases trac
tion increased as tread depth increased. These results agreed with the ear
lier though more limited data presented by Allbert [106].
The data reported in Figures 5.95-5.99 show that velocity has a larger
deleterious effect on thick film wet traction performance than any other
factor, eq. (5.17) implies that traction performance under thick film condi
tions should be inversely proportional to vehicle velocity. This is to be ex
pected since the fluid forces due to both inertia and viscosity increase, and
thus the portions of the interface region in which thick and thin films per
sist will increase as the velocity increases.
There is a complex relationship between load and thick film wet trac
tion performance, so that the net effect that changes in load will have on a
thick film wet traction is not always clear. Table 5.9 summarizes the re
sults of load effect tests conducted by Staughton and Williams [116]. In
general, increases in load are seen to have decreased thick film wet trac
tion at the lowest pressure tested (110 kPa) and increased thick film wet
traction at the highest pressure (330 kPa). In these instances where there

TABLE 5.8 Observed and calculated stopping distance [142]


Wet asphalt road

Rubber Rebound, Hardness Stopping Distance, Stopping Distance,


32° F, (ft.) Calculated (ft.) Observed
BR 62 62 177 180
NR/BR 46 57 162 163
EPDM 32 63 160 160
SBR/BR 33 55 155 158
SBR 34 52 152 151
Butyl 9 • 40 132 130
332 MECHANICS OF PNEUMATIC TIRES

is a reverse in effect with increasing speed, the usual trend is from a de


crease in traction at low speeds to an increase in traction at high speeds.
Home and Joyner [33] presented results for 20 X 4.4 rib tread aircraft tires
and found little load effect on thick film wet traction performance on both
rough textured and smooth concrete surfaces at the inflation pressure of
1007 kPa.
Regardless of the mode of operation, whether it be braking, accelera
tion, or cornering, the presence of thick water films on the pavement sur
face can drastically reduce the available traction. The traction ellipses ob
tained by Gengenbach [112] and illustrated in Fig. 5. 100 show that a large
decrease in traction coefficient occurs as water depth and speed increase.
As discussed previously, wheel slip has a significant effect on wet traction
values, peak values of traction measured at limited amounts of slip often
being considerably higher than locked-wheel values. The complexity of
the overall phenomenon is revealed in data presented by Nordstrom [124]
which show that as either water depth or speed is increased, not only can
shifts occur in the relative traction performance, in peak or in locked-
wheel modes of different tires, but also tires can assume different rankings
in the different modes. These shifts in performance are due in large part to
differences in tread pattern design as was shown by Yeager [7]. His dis
cussion of tread pattern effects on free rolling and locked-wheel dynamic
hydroplaning implies that a block patterned tread design with no longitu-

TABLE 5.9 Effect ofload increase on thickfilm wet traction

Inflation
Tread Pressure Thick Film Wet Traction
Tire Type Condition (kPa) (V - speed in Km/h)
Radial Full Depth 110 Decreased for V < 80
Radial Full Depth 215 Little change for 20 < V < 50
increased for V > 50
Radial Full Depth 330 Little change for V < 70
increased for V = 80
Radial Full Depth 110 Decreased for V < 80
(#2)
Radial Full Depth 215 Little change for V < 30
(#2) increased for V > 30
Radial Full Depth 330 Decreased for V < 40
(#2) increased for V > 40
Crossply Full Depth 110 Decreased for V < 60
little change for V > 60
Crossply Full Depth 215 Little change
Crossply Full Depth 330 Increased
Crossply Worm Smooth 1 10 Little change
Crossply Worn Smooth 215 Little change
Crossply Worn Smooth 330 Increased for V < 60
no change for V » 60
Water Depth - 9.0mm
Surface Smooth Concrete
Based on Data from FIGS. 15-21, 34, 35 of Ref. [116].
CONTACT BETWEEN THE TIRE AND ROADWAY 333

I.OOr

0.75

0.50

g 0.25 I

1.0 Q75 0.5 0.25 0 025 0.5 075 1.0


/IB COEFFICIENT OF FRICTION MA
Braking Driving

FIGURE 5.100. Curves of the coefficient offriction for different velocities and water depths
1112}.
Tire 5.60- 1 5 Load - 250kp Tread Condition 100% P - I.Sail

dinal grooves, having a high peak value of thick film wet traction, will
have much poorer performance under locked-wheel conditions.
As with hydroplaning, it is through its surface texture, porosity, and ag
gregate wear characteristics that the pavement has its principal effects on
thick film wet traction.
Both pavement macro-texture and micro-texture have important roles
in thick film wet traction production. It is through the channels between
elements of macro-texture and through the groove network in the tread
pattern that the bulk of the fluid can be removed from the tire-pavement
interface. Because fluid inertial forces are proportional to the amount of
unchanneled fluid and to the square of the vehicle velocity, how well the
macro-texture functions can have a great effect on how rapidly the thick
film wet traction value will change with speed.
Table 5.10, due to Gough [143], shows how the type of macro-texture
interacts with other factors to create safe and unsafe operating conditions.
The effect of macro-texture on thick film wet traction for different speeds,
water depths, and tread depths is clearly displayed in results obtained by
Nordstrom [124] shown in Fig. 5.101 and by Staughton and Williams
[116] shown in Fig. 5.102.
Because of the importance of macrotexture, numerous experimental de
vices have been developed in an attempt to link any or all of its features to
skid resistance. Over 26 methods have been proposed [144]; however, in
334 MECHANICS OF PNEUMATIC TIRES

TABLE 5.10 Influence of tyre and road factors on safety


(based on data in Fig. 2 of [142])

Micro-Texture of Road Surface


Sharp Polished or
or Harsh Intrinsically Smooth
Best Worst Best Worst
30 30 60 30 60 30 60
Patt
Locked
Open Smooth
or
Coarse Patt
Rolling
Smooth
Patt
Locked
Fine Smooth
or
Closed Patt
Rolling
Smooth
v/ Safe > 0.5
x Unsafe < 0.5

TABLE 5. 1 1 Summary table showing the effect of


road texture on safe operation (number of checks
from Table 5.10)
Micro
Texture
re coefficient > 0.5 Harsh Polished
Macro- Open 11 0
Texture Closed 8 4

all but a few little correlation has been established between the quantities
measured and wet traction values. In fact, two of the most popular cate
gories of measurement techniques, these being volumetric methods such
as the sandpatch [145], grease smear, and silicone putty and drainage mea
surements [145, 146] such as static, pressurized and dynamic drainage,
have been shown to be of at best limited use in predicting wet traction per
formance.
A few researchers have been successful in relating particular measures
of macro-texture to wet traction performance. The size, geometry, and dis
tribution of elements of the macro-texture have been related to the skid re
sistance of the pavement by Schonfeld [147] whose photo-interpretation
method for measuring skid resistance has proved to be useful.
Veith and Hot linger [140] have demonstrated that the traction on thick
CONTACT BETWEEN THE TIRE AND ROADWAY 335

BIAS PLY TIRE ASTM TIRE

02468 02468
WATER DEPTH (mm) WATER DEPTH (mm)
^^ Rough Surface
o a a
931 Omm pattern
IIIIIIIH Smooth Surface » • • » depth
IIUIIU (0-2 mm aggregate)

FIGURE 5. 101 . Effect of water depth, road surface and tire pattern depth on 13% slip braking
force coefficient [124].

water films is improved when the water drainage capacity of the pavement
macro-texture is improved. They have proposed the concept of opera
tional severity as a measure of the tendency for traction loss, where

Operational Severity = a (5.18)


DPHG

where V is the vehicle speed, DP is the dynamic drainage capacity of


drainage paths formed by the pavement macro-texture, //„ is the water
depth on the pavement, //,, is the depth of the tread grooves, and α is a
factor which includes the effect of tread pattern geometry, carcass con
struction, and tread material. Generally conditions leading to a large value
of eq. 5.18 cause a low skid number.
Recently, tire noise has been shown, in an exploratory study [148], to
have great promise as a useful measure of pavement macrotexture in
terms of its effect on wet traction performance.
The contribution of macro-texture to the hysteretic component of tire
traction in the thin film and semi-dry portions of the contact patch is dis
cussed in section 5.7.3 on thin film wet traction.
Pavement micro-texture helps improve thick film wet traction perform
ance principally through assisting the break down and escape of the thin
336 MECHANICS OF PNEUMATIC TIRES

Texture
Depth

o o Quortzite macadam 24mm


» « Rolled asphalt I.Smm
• • Smooth concrete 0.4mm

N\\\\ Influence of surface

Influence of tread
pattern on smooth concrete

New patterned radial


ply tire

Smooth tire on concrete

0 2 4 6 8 10
WATER DEPTH, mm

FIGURE 5.102. Effect of surface and water depth on locked wheel braking force coefficient,
patterned tire; smooth tire results on concrete surface shown dotted [116].

viscous films of water persisting between tread element and pavement sur
faces after the bulk of the fluid has been removed by the drainage chan
nels formed by tread grooves and elements of the pavement macro-tex
ture. It is, thus, the pavement micro-texture in combination with the tire
tread compound and the surface dimensions of individual tread elements
that determines the level of tire traction at low speeds.
5.73 Thin Water Film Effects
When insufficient fluid is present to fill the grooves in the tire tread pat
tern, thin film wet traction conditions exist. Here each element of the tread
pattern must act independently in breaking down the fluid film to estab
lish contact with the pavement surface.
Thin film wet traction performance is a function of many factors, all of
which have been discussed earlier in this chapter. However, certain factors
have a much less important role in thin film wet traction than in the case
of thick film wet traction. These less important factors are tire carcass con
struction, tread depth and inflation pressure, pavement macro-texture, and
fluid density and film depth. References [13, 14, 26, 1 12, 1 13, 124, 138] are
important sources of information in this area.
Considering tire factors first, the construction variables most affecting
thin film wet traction are those which determine the dimensions of and to
CONTACT BETWEEN THE TIRE AND ROADWAY 337

a lesser degree those that determine the pressure distribution in the tire
contact patch. The greater the length of the footprint region and the
higher the pressure levels under the tread elements, the greater will be the
"dry" portion of the footprint and thus the greater will be the thin film wet
traction values [13, 14]. Meades [141] reports a small intrinsic advantage in
peak values for radial tires, which advantage can be related to the higher
mid-footprint pressure levels and reduced squirming action of the tread
elements associated with this type of tire. Radial tires can also accept
smaller tread elements and a greater number of full depth kerfs venting
into the major groove network, both of which aid thin film wet traction
performance [14, 106, 113].
Tread pattern designs effective under thin film conditions have closely
spaced grooves and slots or kerfs. Such patterns reduce the effective flow
distance of any fluid present in the interface and promote the rapid forma
tion of regions of "dry" traction producing contact. In Fig. 5.103 Allbert
shows a strong interrelationship between effective flow distance and limit
ing cornering coefficient. In addition to being closely spaced, these grooves
and kerfs must have sufficient cross sectional area and depth to hold all
water entering the footprint. Gengenbach [112] and Allbert and Walker
[8] show in Figs. 5.106 and 5.104 that for a given water depth and vehicle
speed, there exists a certain critical value of groove area above which thin
film conditions exist and above which traction remains nearly constant.
Kerfs have the additional function of providing sharp edges intensifying
pressure for better film penetration on smooth surfaces and these can also
serve as drainage paths for interfacial pockets of fluid if they open into ad-

„.-._- GROOVE FLOW *


RIBS WIDTH G ROOVES W|DTH DISTANCE
No. (in) No. (in.) (in.)
13 0.36 12 0.10 0.18'
9 0.50 8 0.15 025 Smooth
025 I 7 0.66 6 0.20 0.33 Patterns
5 0.92 4 0.30 0.46.
5 0.66 4 0.30 0.10 Multi-Slotted
5 0.66 4 0.30 0.05^ Patterns
0.20
LJ
u "EFFECTIVE FLOW DISTANCE
\ 1 . Smooth Rib Patterns < Rib Width/2
\\ 2. Multi-slotted Patterns • Distance Between
00.15
Slots/2
o \
z \>D CORNERING FORCE MACHINE TEST
V4 Load: 870 Ib
\<2r Inflation Pressure: 26p.si.
010
185-15 Automobile Tires
\
L

Q05 \
MV
'
0.1 0.2 0.3 0.4 05 0.6
EFFECTIVE FLOW DISTANCE, in.

FIGURE 5.103 Effect of tread pattern "flow distance" on limiting cornering coefficient [106].
338 MECHANICS OF PNEUMATIC TIRES

jaccnt grooves. In fact, on polished surfaces lacking microtexture, only


heavily kerfed tread patterns will have significant levels of wet traction [8,
124].
When design constraints do not permit the use of kerfs in tread patterns,
or when the tire is worn to the base of the kerfs, then the absence of pave
ment micro-texture can lead to situations where the tread block or element
cannot penetrate the thin fluid film. This condition is called viscous hydro
planing, and occurs only on wet surfaces with little micro-texture. Here a
thin film of fluid, usually less than 0.2 mm, persists between tire and pave
ment surfaces because there is insufficient microtexture to force its break
down and allow small pockets of fluid to escape. Viscous forces dominate
in supporting the tire surface in the case of such thin films. The greater the
fluid viscosity, the greater will be the resistance of the fluid to being
squeezed out from between the two surfaces. Viscous hydroplaning can
occur at any speed and with any fluid film depth. However, tire inflation
pressure and water depth are both important in determining whether the
viscous form of hydroplaning will persist at moderate to high speeds—
smaller fluid depths and tires with higher inflation pressures increase the
probability of high speed viscous rather than dynamic hydroplaning.
As in thick film hydroplaning, the forces which the tire can transmit
while in viscous hydroplaning are negligible compared to those which can
be transmitted when the fluid film is broken. Thus, the presence of good
microtexture is an important design objective in highway and runway ma
terial selection and design.
Hydroplaning accompanied by tire tread reversion only occurs when
heavily loaded tires such as those on trucks and aircraft lock up at high
speeds on wet pavements with little microtexture. In addition to the negli
gible values of traction common to other forms of hydroplaning, it is char
acterized by overheating and a resulting "reversion" of the tread rubber

GROOVE GROOVES RIB RIBS


WIDTH. in No. No.
0.02
0.10
0.20
030
0.40

0.2in,Q3ia8
0.4 in. Grooves

0.) in Grooves

0 02 in Grooves
Smooth

20 30 40 50 60 70
SPEED, mph

FIGURE 5.104 Effects of change in groove width on peak braking force (SJ.
Test on experimentally produced five-rib pattern tire: tire, 185 x IS radial ply; water depth. 0-03-0 -08 in; surface,
smooth concrete.
CONTACT BETWEEN THE TIRE AND ROADWAY 339
t.O

0 1 02 0.4 0.6 0.8


1mm TREAD DEPTH, h/h0

FIGURE 5.105 The maximum coefficient offriction of the sideforce as afunction of the tread
depth for different water depths and velocities [112],
Tire 155* Q-y>Okp
Groove Width 3.5mm p - I 5alu

V. on or near the tread surface [106]. Several different theories [6, 108, 109]
have been proposed to explain this condition which, Allbert [106] has sug
gested, may really be two different phenomena having the common end-
point of rubber reversion accompanying low friction, but having different
causes. He noted that in some instances the tread rubber surface showed
signs of melting while in others overheating occurred just below the tread
surface, sometimes with and sometimes without surface melting, causing a
thin film of rubber to be detached from the tread surface [110]. Home et
al, [6] felt that when the surface melts the soft reverted rubber may pro
vide better scaling around the edge of the footprint, thus trapping a thin
film of water which can be heated and changed to steam. This steam pres
sure supports the tire and tends to lift it clear of the runway over most of
the contact patch. No detailed explanation was given for the mechanism
for heating the tread rubber or the water film. Nybakken, et al. [108] of
fered the following step by step theory for the surface-melting type of re
verted rubber hydroplaning. First, a momentary locking of brakes could
cause a rapid temperature rise on the tread surface in the contact patch.
The tread rubber would heat until it became soft and sticky. In the ab
sence of microtexture, and in the presence of at most a moderate amount
of polished macrotexture, the soft tread rubber would conform readily to
the pavement surface, allowing a thin film of water to persist between the
340 MECHANICS OF PNEUMATIC TIRES

0 O.I 0.2 0.3 0.4


GROOVE AREA RATIO

Water Depth 0.2, (0,2. Omm

0 0.) 0.2 0.3 0.4


GROOVE AREA RATIO

FIGURE 5.106. The maximum coefficient of friction of the side force as a function of the
groove area ratio for different velocities and water depths [112].

two. This would explain the low values of traction. The adhesive friction
mechanism for surface heating apparently assumed by these authors does
not account for the damage found below the surface in the second type of
reverted rubber hydroplaning. However, Schallamach's [109] hypothesis
of hysteretic heating below the tread surface due to sliding over wetted
polished pavement macrotexture is a possible explanation for this second
type. Accepting both theories, one would expect the first type of reverted
rubber friction to occur on smooth surfaces, and the second on surfaces
with a moderate amount of polished macrotexture.
Smooth surfaced tires can have large traction values on damp pave
ments as long as significant amounts of micro-texture are present.
These small sharp points in the road surface can penetrate a thin fluid
film, but can also penetrate into the rubber tread surface of the tire. Be
cause the surfaces have thin films of oxide layer and water, they are far
from chemically clean and therefore it is not entirely a water film pene
tration. A microroughness of 0.05-0. 1 mm. on top of the macro texture
produces a high resistance to slip. The adhesive friction process consists of
the formation of adhesive bonds at the real area of contact, at the tips of
the hard asperities, and is caused by the normal load. The elastically
stored energy in shear, due to the tangential force, will try to overcome the
surface energy of the hard solid so as to free adhesive bonds. Only the rub
ber molecules forming the real area of contact may be considered near
enough to the field of forces of the hard sol id and the deformation is there
fore concentrated in a very thin layer below the surface. Under the action
of the tangential force the adhesive bonds break and a fresh cycle then be
gins with formation of new bonds elsewhere on the surface.
In extreme cases, the stresses at the tips will be large enough to rupture
the rubber, causing abrasion, as has been proved in sliding on dry quar
CONTACT BETWEEN THE TIRE AND ROADWAY 341

tzite [93]. The measurement and classification of the road surface micro-
roughness with a mechanical roughness meter is a difficult subject because
the microroughness is superimposed on the texture.
Regression equations from the vehicle tests produced an equation
Skid distance (ft.) = 9.16 + 0.61 R32 + 0.77 D
where /?32 is the Bashore resilience (ASTM D-1054) at 32° F and D is
the Shore A hardness of the rubber.
Stopping distances in Table 5.10 are calculated with this equation. They
show also that skid resistance on wet pavement can be increased in general
by making changes in the rubber, but these usually decrease wear life.
Increases in inflation pressure, according to Browne, et al [13] should
have little effect on thin film wet traction, and this has been confirmed by
Gengenbach [112].

5.7.4 Analytical Models for Fluid Film Effects


A mathematical model for predicting the effect of tread pattern design
on thick film wet traction performance has been the goal of numerous re
search efforts. Grime and Giles [149] found that the ratio D/vCT could be
used as an approximate indicator of the locked-wheel wet traction per
formance of a proposed tread design, where D is the perimeter of the con
tact area A between tire and pavement. Noting the spread in their data,
they suggested that this indicator should serve only as a first approach for
assessing the drainage afforded by different tread pattern designs. Gen
genbach [112] found for circumferentially grooved tires that the peak side
force coefficient was related to the "groove area ratio", this being the ratio
of the area of the grooves in the footprint to the area of the footprint. The
data plotted in Fig. 5.106 shows that this ratio takes a linear form for mod
erate to high speed thick film conditions. Yeager and Tuttie [17] created a
tread pattern parameter called the unit groove capacity, which is found to
relate directly to the thick film wet traction performance and inversely to
the dynamic hydroplaning speed of a free rolling passenger tire. The unit
groove capacity (UGC) is found from the following equation.

-tw,* (5.19)

where W, is equal to the width of the footprint at its center under dynamic
conditions minus the width of one shoulder rib, n is the number of circum
ferential grooves, wi is the average width of the ith groove taken per
pendicular to the longitudinal footprint centerline at the center of the foot
print, and di is the depth of the ith groove. The factor K1 is equal to the
ratio of gross to net contact area, the circumferential groove void area
being excluded from the gross as well as net contact areas.
Veith and Pottinger [140] in addition to their term 'operational severity'
(Eq. (5.18) also developed the concept of "Ultimate Performance Rating"
(UPR), which is proportional to the area A shown in Fig. 5.107:

(5.20)
dV
342 MECHANICS OF PNEUMATIC TIRES

40 60
V.mph

FIGURE 5.107. The ultimate performance rating, UPR. concept [140].

where /ic is the value of cornering coefficient at 60 mph and dμc/dV is the
slope of the graph of μc as a function of velocity, V. Tires with higher val
ues of UPR should have superior performance under conditions of higher
operational severity, as described in eq. (5.20). Tread pattern effects are
thus evaluated experimentally in this scheme. Two additional quantities,
void and "see through" areas have been frequently used, though with only
limited success, as guides for estimating the relative thick film wet traction
performance of proposed tread designs [136].
All of these models are, however, insensitive to many of the details of
tread pattern geometry, and as their limited success in predicting thick
film wet traction performance indicates, are inadequate as tools for tread
pattern design for improved wet traction performance. Recently, Browne
[136] has presented a mathematical formulation of the flow through the
tread pattern groove network which includes both inertial and viscous
fluid forces and which has been used to predict successfully the effects of
changes in groove network design on thick film wet traction performance.
This analytical model, coded for use on a digital computer, calculates a
Traction Rating (TR) for each tread pattern:

TR(A') = 100 (5.21)

where (A′) represents a certain percentage groove depth, and QIN (A′) is
the volume flow rate of water entering the tire footprint at its leading edge
for a tread depth of A percent. By comparing values of Traction Rating
one can rank tread patterns as to their thick film wet traction performance
under locked-wheel conditions, and for the case of smooth or longitudi
nally grooved patterns all values of slip can be considered. Practical appli
cation of the programmed analysis has been in isolating the effects of
changes in single features of the tire tread pattern groove network, such as
the angle of lateral grooves on thick film wet traction as shown in Fig.
5.108. In this way guidelines have been developed for designing most of
the common features found in the tread pattern groove network.
Bathelt [137] has modeled the action under thick film conditions of a
CONTACT BETWEEN THE TIRE AND ROADWAY 343

single tread element of a freely rolling tire with either a straight rib or
block tread pattern. Conclusions are presented on the relationship of
block or rib size to adjacent groove width for good wet traction perform
ance.
A simple static laboratory test has been proposed for evaluating the ef
fect of tire tread pattern design on thick film wet traction performance.
This involves the Delft Tire Tread Drainage Meter [150] where traction
performance is predicted on the basis of the volume of water that can be
forced under pressure (without lifting the tire surface) through the tread
patterns of full scale tires.
No analytical model currently exists for determining the effect of
changes in tire carcass construction on thick film wet traction perform
ance. Two empirical models have been used, both of which involve taking
measurements from full scale tires. Okamura [25] pressed thin rods against
the tread surface of a tire through holes in a flat plate against which the
tire was loaded in order to determine an array of influence coefficients for
the deformation of the tire structure. This array of influence coefficients
was used along with pressures obtained from a purely viscous model for
the fluid flow to determine vertical tire deformation and thus the "wet and
dry" portions of the contact patch. Changes in tire construction are re
flected in changes in the array of influence coefficients. Unfortunately, in
its present form this method is more complex and subject to greater error
than full scale traction tests. Browne [10] used an optical technique, the
moire fringe method, to determine the deformation of the tire surface as it
passed over a water covered glass plate embedded in the pavement sur
face. This deformation, which is different for tires of different construc
tions was used in an analytical model of the fluid flow in the tire footprint
to calculate the effect of changes in tire carcass construction on thick film
wet traction performance.
GROOVE
DEPTH

•s • toon
15.00

^-* »l ao
>/.
0 10.00
i // ^» / "•( 54.0

/ . 1—»
1
*: •00
1
GROOVE
§ ' 16.0 CONFIGURATION
/
• Longitudinal plus
2 5.00' Lateral Grooves
•^ ao • Longitudinal
f Grooves Only

ooo
0 5O 100 150
LATERAL GROOVE ANGLE («)

FIGURE 5.108. Traction rating vs lateral groove angle 19.05 x 15.24 cm tire footprint; two
equally spaced, 1.27 cm wide longitudinal grooves; side only lateral grooves [136].
344 MECHANICS OF PNEUMATIC TIRES

Numerous models have been proposed for the effect of pavement


macro-texture on thick film wet traction. Most of these combine macro-
texture and tread pattern effects, and have been previously discussed [139,
140, 147].
Analytical models have been proposed by Moore [132], Bond, Lees, and
Williams [118], and Bathelt [137] for the effect of macro-texture on the re
moval of thick fluid films from the tire-pavement interface. Good quan
titative agreement has not been obtained between the results predicted by
these models and measured values of thick film wet traction. Analytical
models have been proposed by Tabor [151], Gujrati and Ludema [35], and
Yandel [152] for the contribution of macro-texture to the hysteretic com
ponent of tire traction in the thin film and semi-dry portions of the contact
patch. In general, qualitative but not quantitative agreement exists be
tween the trends predicted by these models and actual thick film wet trac
tion data.
Theoretical analysis of pneumatic tire hydroplaning has presented re
searchers with some formidable problems. Unless one is considering the
special case of purely viscous hydroplaning, both viscous and inertia! ef
fects must be included in the analysis of the fluid flow. In addition, flexi
bility of a surface bounding a pressurized fluid has been found in lubrica
tion studies to have a dramatic effect on the nature of the fluid flow [134].
Thus an analytical model should incorporate tread element flexibility in
the case of viscous hydroplaning and tire carcass flexibility in the case of
dynamic hydroplaning. Finally, surface texture, has been found experi
mentally to have a strong influence on the degree of hydroplaning, and it
should be considered at some point in a comprehensive model of the phe
nomenon.
Currently the analytical model of Browne, et al [135] is the most com
plete description of pneumatic tire hydroplaning in those instances where
sufficient fluid is present to flood the tread pattern groove network. Both
viscous and inertial effects are included and the possible effect of turbu
lence is analyzed. Carcass flexibility is introduced through the use of an
experimentally determined film thickness distribution under the tire [10].
This approach is necessary since there exists no analytical model of the
tire structure sufficiently accurate to supply such data. The pavement sur
face was assumed to be smooth to allow a steady state formulation of the
problem. In general, most analytical formulations of the effect of surface
texture have been confined to studies of the sliding of rubber over a single
lubricated asperity. An example of an attempt to include surface texture in
a global sense in the case of viscous hydroplaning is the work of Daug-
haday and Tung [24]. The formulation by Browne has proved to be a pow
erful tool in analyzing the hydroplaning phenomenon. With its help, tire
carcass deformation along the centerline of the footprint has been identi
fied as the chief factor allowing dynamic hydroplaning to occur at rela
tively low speeds. Detailed views of the fluid flow distribution in the foot
print region obtained with the model have helped to determine the effect
of individual tread design features on the fluid flow through the footprint.
This, along with additional information supplied by the analysis, allows
one to determine how effectively fluid would be channeled by a groove
network design and to determine hydroplaning characteristics of proposed
tread patterns [136].
Modeling of the tire-road interaction in the case of thin films of water
requires a departure from the concepts of classical hydrodynamics. Atten
CONTACT BETWEEN THE TIRE AND ROADWAY 345

tion must be focused on local contact pressures and traction stresses, and
this includes the consideration of several modes or components of sliding
resistance. On the assumption that a road surface is a carpet of discrete
bumps, spherical in shape, Gujrati and Ludema calculated the drag force
of a single sphere sliding on rubber in the presence of a lubricant. The
analysis used the equations of elasto-hydrodynamics modified to accom
modate the visco-elastic properties of rubber. The properties of the rubber
are obtained from an indentation test. The most important consequence of
the visco-elastic properties of rubber are that the contact between the
sphere and rubber is asymmetric, chiefly because the rubber recovers
slowly in the rear portion of contact. A further consequence is that the
fluid film thickness will be less than predicted from symmetric contact,
which produces higher drag force.
The analysis correlates well with experiments using water as the lubri
cant except at low speeds. Experiments using fluids that wet rubber better
than does water show good correlation at very low sliding speeds as well.
The authors suggest that practical tire traction may well be influenced by
the poor wettability of water to rubber.
Gujrati and Ludema further suggest that where the strain fields in rub
ber interfere with each other due to closely spaced spheres, the drag force
may be affected. Yandell [152] shows that the effect of closely spaced in-
dentors is in fact to decrease the hysteresis factor in sliding friction. He at
tributes this reduction to a "stress saturation" of the rubber although his
calculations are done on the basis of a fixed spacing between indentors,
but with increasing applied load.
Besides this model of traction due to the interaction between tread rub
ber and pavement macro-texture, several additional empirical formulae
and analytical models have been proposed as descriptions of the phenom
ena associated with traction under thin film conditions. Before entering
into a discussion of these models, however, it is important to establish
some criteria for determining when thin film conditions exist. Sinnamon
and Tielking [113] offered the following dimensionless ratio, denoted by
γ1, as a guideline for separating thick and thin film conditions:

y,= = (5-22)

where V is the rate at which groove volume becomes available to absorb


water, () is the volume rate of water in the tire path, g is the groove vol
ume per unit of gross contact area, and h is the depth of water on the pave
ment measured from the top of asperities. A value of γ , greater than four
was found to be a good indicator of the existence of thin film conditions in
tests in which peak braking forces are being measured; a much larger,
though unspecified value of y, was considered necessary for thin film con
ditions to persist during locked-wheel braking.
Grime and Giles [149] and Allbert and Walker [8] have offered empiri
cal formulae found to be in moderate correlation with locked-wheel brak
ing coefficients, fiB, as in eq. 5.23 and limiting cornering coefficient, ,u , eq.
5.24 respectively:
. perimeter of contact area .. „..
M* = k -- :- (5.23)
Vcontact area
346 MECHANICS OF PNEUMATIC TIRES

and

(5.24)
average distance between grooves or slots if present

where k is a constant dependent on the pavement surface and N is approx


imately one. Sinnamon and Tielking [113] provided an approximate ana
lytical interpretation for such flow distance parameters. Their analysis
showed that under thin film conditions, γ2, a parameter defined in eq.
(5.25), should be an indicator of the difficulty in establishing contact be
tween a tread element of a particular geometry and the pavement surface:

*- (5'25)

where A is the actual contact area between tire and road, C is the perime
ter of this area, and L is the length of the contact.
In thin film hydroplaning, where there is insufficient fluid to fill the
groove network, each tread element acts independently in its attempt to
penetrate the fluid film and establish contact with the pavement surface.
Several analytical models have been proposed for this phenomenon. Saal
[129] first applied the classical rigid body squeeze film model to the prob
lem. Hays [130] extended the analysis to rectangular tread elements.
Moore [131, 132] approximated the effects of tread element inclination-ro
tation and pavement surface texture. More recently Bathe [137] presented
a detailed model for the sinking action of a rigid tread element of a freely
rolling tire through a film sufficiently thick (>0.2mm) so that fluid inertial
effects were important. Pavement macrotexture and surface drainage
channels were included in an approximate manner and were found to
have a dominant effect on eliminating hydroplaning under such condi
tions. Browne, Whicker, and Rohde have performed a series of studies [13,
14, 138] in which the squeeze film model was reformulated to include the
important effect of tread element flexibility. The effects of tread element
rotation and pavement surface texture were included in a more rigorous
manner.
The work of these authors has shown that traditional rigid body squeeze
film analysis is not an accurate model for viscous hydroplaning, but that
viscous hydroplaning occurs when the bell-shaped pockets of fluid trapped
under the flexible elements upon entering the footprint cannot escape due
to a lack of adequate pavement microtexture. Kerfs which vent into adja
cent grooves would help drain fluid trapped in such pockets, and in this
manner would aid in preventing viscous hydroplaning. The recent series
of studies by Browne, Whicker, and Rohde [13, 14, 26, 136] has clarified
the importance of many factors previously either ignored or inadequately
treated. Figure 5.109 illustrates the basic geometry used in their analyses.
The Reynolds equation for the intervening fluid film:

V—VP-*. (5.26)

is modified in the sense that the film thickness h(X,Z,T) between tread ele-
CONTACT BETWEEN THE TIRE AND ROADWAY 347

t,

TOP VIEW T
L

1
Initial position of
backing plate

Rigid backing plate

-Flexible treod element


' Initial position of
tread surface

Smooth Pavement

FIGURE 5.109 Flexible tread element squeeze film geometry [13]

ment and pavement is expressed as a function of fluid pressure P(X,Z,T)


to account for the flexibility of the tread element:
h(X,Z,T) = h0(T) + LP (5.27)
Here L represents a linear operator transforming the surface loading into
tread element deformation, h0(T) is the tread element surface reference
level at time T, and /i is the fluid viscosity. Slip or relative velocity U be
tween tread element and pavement surfaces was allowed through the addi
tion of the term U/2 ∂h/∂X to the right hand side of eq. 5.26. Rotation of
the tread element as shown in Fig. 5.1 10 was included by expressing h(,(T)
in the following manner
h0(T) = //„(/) - D + Xtan9(T) (5.28)

Initial position of
backing plate

\ a =8(0) Instantaneous
I ,- / position of
backing plate

FIGURE 5. 1 10 Motion ofsinkage element [138]


348 MECHANICS OF PNEUMATIC TIRES
IA
Hoys - Rigid tread element
_ I I I
1.0 hM(N- Browne, Whicker, Rohde -
Flexible tread element
h

OS

\
0.001 0.01 0., _ 10 100 1000

FIGURE S.I 1 1 Time history offluid film thickness n; comparison offlexible with rigid tread
element model; constant load, no slip III = 0); hmla = minimum film thickness [13].

where the time dependence of θ(T) was based on the limited amount of
tire deformation data available [10]. Surface micro-texture in the form of
concentric rings was incorporated in the model for the case of a circular
element by Rohde [26].
Only work on the precontact problem, i.e. sinkage of the tread element
surface until it first touches the pavement, has so far been published but
this has provided several important insights to the problem. Tread element
flexibility was found to have a profound effect on the fundamental nature
of the squeeze film event as shown in Fig. 5.111. Classical rigid plate,

LIQUID nit
-Woter ot28°C 8.59 10 1.319 0.05
-Diesel Fuelot28°C 259 KT1 1.717 0.07
-Oil SAE # 10 at 28°C 6.25 KT1 3636 0.15
Oil SAE #40 at 280C 223 icr1 5.420 0.22
-Woter at 0«C 189 1.601 006
-Diesel Fuel ot 0°C 5.8 id* 2.059 0.08
1.0 -
19.1 mm x 19.1 mm
Tread Element

0.01 100

FIGURE S.112 Time histories of minimum film thickness, nmhvfor different fluid viscosities.
Tc — Time of initial contact; T,,u — time tread element is in footprint [14].
CONTACT BETWEEN THE TIRE AND ROADWAY 349

squeeze film analysis was, thus, shown to be inadequate as a tool for de


signing tread elements for good thin film wet traction performance.

5.8. Tractive Forces on Snow and Ice


5.8.1 Traction on Ice
Ice is most likely to form by the freezing of water already adhered to the
road surface. It will have the effect of progressively obliterating the surface
texture of the road surface beginning with the micro-texture. In the limit it
produces a smooth surface.
The design of tires for ice would appear to be done on the same prin
ciples as the design of tires for smooth mastic asphalt surfaces, and this
seems to apply to very low temperatures. On the other hand, when sliding
occurs, the surface of the ice melts and the resulting water film becomes an
effective lubricant, with little hope of a hysteresis component of friction
because of the smoothness of the ice. The temperature rise at a sliding sur
face may be approximated using the equation originally present by
Barnes, Tabor and Walker [153]:
1
4a

where
A T is the temperature rise
ft. is the coefficient of friction
L is the applied load
g is the gravitational constant
ν is the sliding speed
a is the half width of the tire
k1 and k2 are the thermal conductivities of the ice and the rubber.
Using a value of 0.3 (on dry ice), a slip velocity of .02 mph and automotive
tires, a temperature rise of 25 °F may be expected. Thus at 7°F, slip be
tween a tire and ice may cause some melting, beginning the process of the
reduction of friction. No complete analysis has yet appeared to completely
characterize the water film thickness and properties. It is confirmed how
ever that melting of ice due to contact pressure is not a factor under tires
such as exists under skates. The melting point suppression is calculated to
be approximately 0.0074°C/bar.
Again applying the principles of lubrication to the operation of tires on
ice, one would expect that large contact area would increase viscous drag,
and this would be achieved by low inflation pressure and to some extent
by soft rubber where there is a tread pattern. Water film thickness would
be the greatest where slip is large (wheel lockup) and where tread features
are large, resulting in low friction. On the other hand a tire with a large
contact area and small tread features should be desirable.
Complete testing of all variables influencing tire traction on ice has not
been done. The recent papers of Southern and Walter [154] and Gnovich
[155] discuss the major variables of importance.
350 MECHANICS OF PNEUMATIC TIRES

5.8.2 Traction on Snow


Traction is drastically reduced on snow-covered pavements, often being
quite low even compared to that under wet conditions, and may be re
duced sufficiently to make it impossible to negotiate grades. In addition,
starting values of traction on snow can be so low that a vehicle can be im
mobilized. These are serious problems and literally thousands of snow
traction tests have been run over the years [157]. Because of the extreme
complexity of the problem it has been only within the last few years that
an understanding has emerged of the fundamental mechanisms contrib
uting to snow traction. [158-160].
This discussion of snow traction will begin with an overview of the me
chanics of snow traction in which fundamental mechanisms will be em
phasized. Important findings from snow traction tests will then be re
viewed. Finally a brief discussion will be made of different treatments,
chemical and material, used to raise the skid resistance of snow-covered
pavements.
Mechanics of Snow Traction
Fundamental factors affecting snow traction can be arbitrarily grouped
under three headings: tire parameters, snow properties, and operating con
ditions.
Tire Parameters
The tire tread pattern, tread compound, and carcass construction, are
important to the development of snow traction.
The open tread patterns with relatively large groove areas appearing on
snow tires are designed to increase traction by using the edges of tread
blocks to contact and/or dig through the underlying snow cover. To do
this effectively the grooves must be self cleaning, which is a major reason
for the predominantely wide-grooved designs in use today. Some tread
patterns, however, are designed purposefully to retain snow, with the idea
of increasing traction by the bonding together under pressure of the re
tained snow with that covering the roadway [166]. Traction forces can be
generated by a tire at five different locations of tread-snow or snow-snow
interfaces [159]:
1. outer tread surface
2. longitudinal grooves
3. effective cross grooves
4. outer sidewall edge surfaces
5. sloping leading edge of the tire-snow contact zone
Mathematical formulations for the traction producing mechanisms as
sociated with each of these interface regions were made in ref. [159]. As
an example of such a mechanism, consider the action of a cross groove on
a tire rolling with a small amount of slip, shown in Fig. 5.1 13. Snow is ini
tially compacted under the outer surfaces of the tread elements. Traction
is first produced by the shear of the snow wedges in the cross grooves
along the dotted lines shown in Fig. 5. 113A. As slip progresses, the cross
ribs sink into the less firmly compacted snow that previously lay under the
cross grooves, as in Fig. 5.1 13 B. The ribs next encounter and then shear
the wedges of more firmly packed snow which were formerly under the
CONTACT BETWEEN THE TIRE AND ROADWAY 351

Vv«hicl«

FIGURE 5.113. Digging action of the tread [159].

ribs. This digging action of the tread not only supplies traction directly but
may permit the tire to penetrate the snow cover and establish contact with
the pavement surface.
Other devices are often used to further aid traction. Among these aids
are studs, elastomeric belts and cleats, and metal chains, all of which me
chanically improve digging capabilities of the tire. [161, 162].
Special rubber compounds are used for the treads of winter tires to in
crease their tractive capability on snow as well as on ice. Natural rubber
(NR) and oil extended natural rubber (OENR) provide superior traction
on ice and hard packed snow than styrene-butadiene rubber (SBR) tire
tread compounds [163, 164]. Since the use of studded tires is now prohib
ited in many areas, there is much interest in learning more about the effect
of tread compound on snow and ice traction. Two theories have been ad
vanced for explaining the action of different tread compounds on snow
and ice. The first theory is based on the wettability of tread rubber. [158,
165]. According to this theory hydrophilic tread compounds promote the
adhesion of thin films of water to the tread, and make it more difficult for
contact pressures to expel such films, resulting in lower traction values.
The analysis of Browne, Whicker, and Rohde [13] of the squeezing action
of flexible tread elements on wetted smooth surfaces supports this hypoth
esis. Hydrophobic or non-wetting surfaces allow water films to escape, en
hancing traction between tread and snow or ice.
The second theory of winter tire tread compounds states that traction on
packed snow and ice is related to rubber flexibility—tread rubber that re
mains flexible at low temperatures is able to conform more closely to, and
establish intimate contact with the underlying surface [158]. Traction is as
sumed to increase as flexibility of the tread compound increases. Some of
the methods to achieve this have been to use silicate fillers instead of car
bon black, to include natural rubber in the tread mix, and to treat the
tread rubber with low freezing point mineral oil.
Snow and ice traction performance can often be increased substantially
by the inclusion of additives to the tread formula. Such varied additives as
chopped wire or wire coils, ground walnut shells, oak hulls, and sawdust
have been used, mostly in recapped tires, to increase traction by providing
many gripping edges and voids [167-169, 174]. The rate of tread wear is
usually substantially higher for tires with such compound mixes.
A temporary increase in traction on packed snow or ice can be obtained
by using one of the commercially available sprays or brush-on solutions
that are designed to be applied to the tread surface immediately prior to
use. A more permanent increase in traction on such surfaces is produced
by cutting or sawing numerous small lateral slits or lacerations in the tread
352 MECHANICS OF PNEUMATIC TIRES

surface [167, 170], so that the number of kerf edges is greatly increased.
Locally high pressures at each kerf edge improve the wiping and digging
action of the tread surface.
The carcass construction of a tire, whether it be radial, belted bias, or
bias, can have both direct and indirect effects on snow traction perform
ance [159]. Carcass construction can affect snow traction directly through
the footprint pressure distribution and resulting differential tread element
sinkage, through the dimensions of the contact patch, and through the
squirm action of the tread. However, the major influence which the car
cass construction can have on snow traction performance is through the
wider range of tread patterns permissible on radial tires. The radial con
struction allows a greater number of wider grooves compared to bias tires
while still showing less wear.
Snow Properties
The physical properties of the snow surface determine the nature of the
interaction between tire and snow which produce traction. The moisture
content, degree of compaction and penetrability, and temperature of the
snow cover are of greatest importance. Schaerer [171] found that snow
with an initial free-water content of less than 15 percent usually was rap
idly compacted by traffic, that with an initial free-water content between
15 and 30 percent snow remains on the highway but flows readily with
little shear strength under the action of the tread, and that with an initial
free-water content greater than 30 percent snow is rapidly cleared from
the pavement by traffic. Highly compacted snow cannot be penetrated by
tread pattern elements. Traction on such snow can thus be provided only
by a combination of mechanical interlock and plowing supplied by add
on devices such as chains and studs, by mechanical bonds produced by
special additives such as chopped wire in the tread mix and/or by adhe
sive bonds resulting from the use of soft, hydrophobic tread compounds.
On either lightly compacted or loose snow, the digging and plowing action
of either tread elements or mechanical add-on devices is of prime impor
tance. Temperature can drastically affect snow traction mechanisms and
friction values, especially on hard packed snow. Higher temperatures pro
duce surface water films dramatically reducing traction levels except when
devices such as chains or studs are added or special additives are included
in the tread mix. Higher temperatures on loose snow lead to its more rapid
dispersal by traffic.
In addition to determining the nature of the interactions producing trac
tion between tire and snow, certain physical properties of snow also deter
mine the levels of available traction [159]. Among these properties are the
shear strength of snow, the compression force-displacement characteristic
for snow, and the shear stress-displacement characteristic for snow.
In general, this area of snow properties is ill-defined and should be a
principal focus of future snow traction research.
Operating Conditions
Both the absolute value of snow traction given by a tire and its snow
traction performance relative to that of another tire can change signifi
cantly with changing operating conditions. High traction levels are desired
for starting, accelerating to speed, negotiating grades, cornering and decel
eration. Test procedures have been proposed and used to evaluate snow
traction performance under each of these conditions [157, 161, 167, 172].
CONTACT BETWEEN THE TIRE AND ROADWAY 353

Unfortunately, what has been learned is that the traction contributions of


different mechanisms are differentially affected by different operating con
ditions, so that no single approach will maximize traction under all condi
tions. In general, however, relatively low tire slip produces maximum val
ues of traction on a hard-packed snow surface due to the combined
operation of add on devices, tread mix additives and tread compound ef
fects, while a higher slip rate maximizes traction on lightly packed snow
due to the digging actions of the snow tire tread and add on devices.

5.83 Snow Traction Tests


Recommended procedures for snow traction tests are given in several
references [162, 167, 172]. Due to the large variability in results obtained
on snow, and the alteration of snow properties with repeated testing, much
of the testing of snow tires is done on ice where test conditions can be
more closely monitored.
The principal objective of most snow traction tests is to obtain relative
rankings for the performance of a group of tires. Only a limited number of
studies have been reported in which the intent was to isolate the effect of a
single design element. A brief summary will be made here of snow trac
tion test results with emphasis being placed on the findings of such studies.
Tread Pattern Design
The effect of tread pattern design on snow traction performance is in
variably tested and reported in either of two different ways: a group of
snow tires with different tread designs is tested and then the individual
tires are ordered as to their relative snow traction performance [169], or
the average performance of a group of snow tires is compared to that of a
highway tire or tires [162]. As an example, traction data for 14 brands of
winter tires plus that for a new cross ply tire are shown in Fig. 5.1 14. The
shaded areas represent the range of winter tire traction data. In addition to
the fact that the tread patterns used are so different that one can learn little
about the effect of changes in individual tread pattern features, the effect
of tread pattern design is confounded with those of tread compound and
tire carcass construction.
The effect of an add-on snow traction aid is readily determined by test
ing similar tires with and without such a device. Sapp [167] reports a 2 to
14 percent increase in traction (acceleration without slip) and a 10 to 34
percent increase in locked wheel skid resistance on hard-packed snow due
to adding studs to rear tires. Smith and C lough [164] found an average in
crease in starting traction of six percent with studs, and of 22 percent with
reinforced steel tire chains on hard packed snow. Tires with snow chains
consistently outperformed those without on hard-packed snow at values of
slip exceeding 20 percent as shown in Fig. 5.1 14. Similar results were re
ported in ref. [161]. On loosely packed snow, reinforced tire chains pro
duced a 20 percent decrease in stopping distance, a 100 percent increase in
spinning traction, and a 70 percent increase in starting traction as com
pared to a typical snow tire in tests reported in ref. [169].
Tread Compound
Different tread compounds can be tested on otherwise similar tires to
isolate their effect on snow traction. Test results reported by Reinhart
354 MECHANICS OF PNEUMATIC TIRES

20 40 60 80 100

40 60
% SLIP

FIGURE 5.114. Slip dependence of driving and braking force coefficients on hard-packed
snow at 10 km/hr [1 73].

[168] for a locked- wheel skid from 20 mph on hard-packed snow ranked
tread compounds on the basis of their average skid resistance. Best was
natural rubber, next was 37.5 phr oil extended GR-S, GR-S, and last butyl
rubber. Tires with treads of natural rubber were found to have 6 percent
greater starting traction [162] and superior hill -climbing ability [164] than
tires with synthetic rubber treads on hard-packed snow.
In a similar manner, the effects of various coarse particular tread mix
additives and of tractionizing treatments— sawing or cutting slits in the
tread—can be found. Table 5.12 contains test results obtained by Sapp
for the traction contribution of four different additives on hard-packed
snow. These additives are seen to be more effective in increasing starting
traction than in improving locked-wheel skid values.
In other tests adding 8 cuts per inch, Sapp obtained a 30 percent im

TABLE 5.12 Test results on hard-packed snow. From [167]


Traction Skid
Index Index
X Control 100 100
X + 4% Fiber 120 110
X + Shell 123 113
X + Var. 143 128
X + Chopped Wire 136 113
CONTACT BETWEEN THE TIRE AND ROADWAY 355
provement in starting traction and a 25 percent gain in locked-wheel skid
values, both on hard-packed snow.

Tire Carcass Construction


Carcass construction effects can be isolated by testing tires that are simi
lar in all respects other than construction. Sapp [167] reports an 11-13 per
cent improvement in starting traction on hard-packed snow when switch
ing from bias ply to radial tires but only negligible differences in skid
traction values.

Antiskid Materials
In order to raise the skid resistance of snow and ice covered pavements,
a wide range of paniculate materials and chemical treatments are fre
quently used, either by themselves or after plowing of the pavement. Heg-
mon and Meyer [175] investigated the relative performance of four fre
quently used materials: boilerhouse and coke cinders, sand, and crushed
limestone. Repeatability of outdoor tests was poor so that a special indoor
cold room facility with a circular track was used. Results for tests with
coke cinders—plain and treated with calcium chloride (a standard melting
agent)—on packed snow are shown in Fig. 5.1 15. Tests of the other three
materials produced similar results, indicating equal effectiveness of the
four materials. General conclusions based on these tests were
all four materials produce a significant initial increase in friction.
ii). friction decreases steadily with increasing number of wheel passes due
to the embedding of the materials in the snow or their being dis
lodged from the wheel path.
iii). using a larger initial amount of material will increase the friction level
throughout the test.
iv).fine particles passing a No. 50 sieve contribute little to any improve
ment in friction.
Ichihara and Mizoguchi [161] studied the effectiveness of three different
chemicals, NaCl, CaCl2, and MgCl ,, in raising the skid resistance of snow-
covered pavements. Skid resistance levels were, in general, raised 0.05 to
0.15, no difference being detected between the effectiveness of the differ
ent chlorides.

8 16 24 32
NUMBER OF WHEEL PASSES , n

FIGURE 5.115. Coefficient offriction f vs number of wheel passes n for 0.6 Ib/sg yd of coke
cinders, treated with calcium chloride and untreated, on hard-packed snow [175J.
356 MECHANICS OF PNEUMATIC TIRES

5.9 Tractive Forces on Unpaved Surfaces


In addition to traction, motion resistance due to tire sinkage and the
available drawbar pull (tractive effort in excess of that required for for
ward motion) are important factors in tire design and selection for off-
road locomotion. In addition to operational demands the nature of the un
derlying surface is thus of primary importance to the tire designer.
There exists a wide range of off-road surfaces: fine-grained and coarse
mineral soils, factional and cohesive soils, stratified media, organic soils
such as turf, muskeg, etc., and snow, to mention a few. Bekker [27] and
Bekker and Semonin [191] have developed computational procedures for
determining the major components of motion resistance—the energy con
sumed in vertical soil compaction, the energy spent in horizontal dis
placement of the soil, and the energy lost due to tire carcass deflection—
on the three basic types of terrain— homogeneous soil, stratified soil with a
more easily penetrated layer lying above a more impenetrable one, and
stratified soil in which the layers have just the reverse properties. Bekker
[192-194], Czako [195], Janosi [196], and Satake and Mukai [197] among
others provide expressions for predicting the available drawbar pull on the
different types of surfaces.
In addition to traction, flotation, and drawbar pull, there are myriad
other operational demands that may be placed on the tire-soil pair.
Among these demands which frequently can be met only by tires designed
specifically for operation on a particular type of surface are better resis
tance to penetration and impact fatigue in tread and sidewall regions, bet
ter cut and wear resistance of the tread region, reduced soil compaction or
increased soil compaction, reduced rutting of the ground surface, reduced
damage to surface vegetation, and better cushioning for reduced vehicle
damage and improved driver comfort. [198-200]. However, many tires
such as those on earthmovers, ATV vehicles, and agricultural equipment
are used under a wide range of conditions including different surfaces,
loads, speeds, and lengths of operation. Most such tires cannot be designed
for maximum performance under a single set of operating conditions.
They must be designed to meet a wide variety of often conflicting de
mands if overall adequate performance is to be achieved.
Tires for off-road use are available in numerous sizes, ply ratings, mate
rials, tread patterns, and depths and load and inflation pressure ranges
[201]. Unfortunately, only a few guidelines, such as the general usage rec
ommended by the manufacturer and the structural ratings tabulated in the
TRA Yearbook [202] are available for tire selection. The recently recom
mended ton-mile-per-hour rating system (TMPH) should prove extremely
helpful, allowing users to determine the maximum haulage/work capacity
of different earthmover tires for an intended job [201, 203]. As an ex-

TABLE 5. 13 Ton-Mile-Per-Hour Ratings. From [200J

Nylon Bias Ply Radial Ply


Size E-3 E-4 E-3 E-4
18.00-25 120 100 150 125
27.00-49 285 250 470 300
29.5-29 143 260
33.25-35 210 400 220
CONTACT BETWEEN THE TIRE AND ROADWAY 357

ample, in Table 5.13 TMPH ratings are used to compare the performance
of two types of radial ply and nylon bias ply tires for four different tire
sizes. In this table, higher numbers imply better performance.

References
[1] Hadekel, R., The Mechanical Characteristics of Pneumatic Tyres, S. and T. Memo No.
10/52 (British Ministry of Supply, TPA 3/T1B, 1952).
[2] Michael, F., Zur Frage der Abmessungen von Luftreifen für Flugzeuglaufra'der (The
Problem of Tire Sizes for Aircraft Wheels). Jahrbuch 1932 der D.V.L., 3, p. 17,
Available in English translation as NACA TM 689 (1932).
[3] Dodge, R. N. (Unpublished data).
[4] Stresses Under Moving Vehicles, Report 4 (U.S. Army Engineer Waterways Experi
mental Station Technical Report No. 3-545, July 1964).
[5] Seitz, Norbert, Die Kra'fte in der Bodenberubhrungsflache schnell rollender Reifen,
Fortschritt-Berichte, VDI-Zeitschrift 12 (19), 46 (1968).
[6] Home, W. B. and Dreher, R. C, "Phenomena of Pneumatic Tire Hydroplaning,"
NASA TN D-2056, National Aeronautics and Space Administration, Washington.
D. C, November 1963.
|7] Yeager, R. W., "Tire Hydroplaning: Testing, Analysis, and Design," in The Physics of
Tire Traction, Theory and Experiment, D. F. Hays and A. L. Browne, Eds., Plenum
Press, New York, 1974, pp. 25-64.
[8] Allbert, B. J. and Walker, J. C., 'Tyre to Wet Road Friction at High Speeds," Proc.
Instn. Mech. Engrs., Vol. 180, Pt. 2A, No. 4, 1965-66.
[9] Moore, D. F., "Drainage Criteria for Runway Surface Roughness," J. Roy. Aero. Soc.,
Vol. 69, 1965, pp. 327-342.
[ 10] Browne, A. L., "Tire Deformation During Dynamic Hydroplaning," Tire Science and
Technology, TSTCA, Vol. 3, No. 1, Feb. 1975, pp. 16-28.
[1 1] Roberts, A. D., "Squeeze Films Between Rubber and Glass," J. Phys. D.: Appl. Phys.,
Vol. 4, 1971, pp. 423-435.
[12] Field, G. J. and Nau, B. S., "Lubricant Behavior in Loaded Rubber Contacts," Wear,
Vol. 35, 1975, pp. 79-85.
[13] Browne, A. L., Whicker, D., and Rohde, S. M., "The Significance of Tread Element
Flexibility to Thin Film Wet Traction," Tire Science and Technology, TSTCA, Vol.
3, No. 4, Nov. 1975, pp. 215-234.
[14] Browne, A. L. and Whicker, D., "Design of Tire Tread Elements for Optimum Thin
Film Wet Traction," SAE paper no. 770278.
[15] Karafiath, L. L., "Soil-Tire Model for the Analysis of Off-Road Tire Performance,"
Grumman Research Department Memorandum RM-541, May 1972.
[16] Freitag, D. R . Green, A. J., and Murphy, N. R., Jr., "Normal Stresses at the Tire-Soil
Interface in Yielding Soils," Miscellaneous Paper No. 4-629, U.S. Army Engineer
Waterways Experiment Station, February 1964.
[17] Yeager, R. W. and Tuttle, J. L., "Testing and Analysis of Tire Hydroplaning," SAE
Paper No. 720471, 1972.
[18] Smiley, R. S., and Home, W. B., Mechanical Properties of Pneumatic Tires with Spe
cial Reference to Modern Aircraft Tires, NACA Technical Note 4110 (Jan. 1958).
[19] Whittemore, H. L., and Petrenko, S. N. Friction and Carrying Capacity of Ball and
Roller Bearings, Nat. Bur. Stand. (U.S.), Tech. Note 201 (1921).
[20] Böhm, F., Mechanik des Giirtelreifens, Ingeniur-Archiv 25, 82 (1966).
[21] Fiala, E., and Willumeit, H. P., Radiale Schwingungen von Gürtel-Radialreifen, A. T.
Z., 68 (2), 33 (Feb. 1966).
[22] Clark, S. K., The rolling tire under load, SAE Paper No. 650493 (SAE Midyear Meet
ing, Chicago, 111., May 17-21, 1965).
[23] Dodge, R. N., Prediction of pneumatic tire characteristics from a cylindrical shell
model. Report 02957-25-T, (Office of Research Administration, The University of
Michigan, Ann Arbor, Mich., March 1966).
[24] Daughaday, H. and Tung, C., "A Mathematical Analysis of Hydroplaning Phenom
ena," Cornell Aeronautical Laboratory Technical Report No. AG-2495-S-1, Buf
falo, Jan. 1969.
[25] Okamura, H., "A Research on Hydroplaning Phenomena in Consideration of Tire De
formation," Japan Automobile Research Institute Technical Report No. 14, August
1974.
[26] Rohde, S. M., "On the Combined Effects of Tread Element Flexibility and Pavement
Microtexture on Thin Film Wet Traction." SAE paper no. 770277.
358 MECHANICS OF PNEUMATIC TIRES

(27] Bekker, M. G., "Practical Aid to Off-Road Tire Evaluation with Bevameter Tech
niques," SAE paper no. 741133, 1974.
[28] Kraft, P., Force distribution in the contact surface between tire and runway, NACA
TM 1365 (Aug. 1954).
[29] Markwick, A. H. i > . and Starks, H. J. II. Stresses between tire and road," J. Inst. Civil
Engineers 16, 309 (1941).
[30] Bode, G., K räfte und Bewegungen unter rollenden Lastwagenreifen, A. T. Z. 64 (10),
300-306(1962).
[31] Lippmann, S. A. and Oblizajek, K. L., "The Distributions of Stress Between the Tread
and the Road for Freely Rolling Tires" SAE paper 740072, 1974.
[32] Tret'yakov, O. B. and Novopol'skii, V. I., "Distribution of Contact Stresses over the
Projections of Treat Patterns," Soviet Rubber Technology, Vol. 28, No. 8, August
1969, pp. 40-43.
[33] Home, W. B. and Joyner, U. T., "Pneumatic Tire Hydroplaning and Some Effects on
Vehicle Performance," SAE Paper No. 970C, Jan. 1965.
[34] Giles, C. G. and Sabey, B. E., "The Effect of Pressure and Friction on Photographic
Emulsions," British Journal of Applied Physics, Vol. 2, No. 6, 1951, p. 174.
[35] Gujrati, B. D. and Ludema, K. C., "An Analysis of Some Factors that Influence Wet
Skid Resistance," The Physics of Tire Traction, Theory and Experiment, Eds. D. F.
Hays and A. L. Browne, Plenum Press, New York, 1974, pp. 197-212.
[36] Yandell, W. O., "A New Theory of Hysteretic Sliding Friction," Wear, Vol. 17, 1971,
pp. 229-244.
[37] Hegmon, R. R., 'The Contribution of Deformation Losses to Rubber Friction," Rub
ber Chemistry and Technology, Vol. 42, 1969, pp. 1122-1135.
[38] Schallamach, A., "Friction and Frictional Rise of Wedge Sliders on Rubber," Wear,
Vol. 13, 1969, pp. 13-25.
[39] Moore, D. F., "An Elastohydrodynamic Theory of Tire Skidding, FISITA, Twelfth
International Automobile Technical Congress Proceedings, Report 2-02, 1968.
[40] Moore, D. F., The Friction and Lubrication of Elastomers, Pergamon Press, New York,
1972.
[41] Green, A. J., Normal Stresses at the tire-soil interface in yielding soils (U.S. Army Wa
terways Experiment Station Report, Vicksburg, Miss.).
[42] Freitag, D. R., and Green, A. J. Distribution of stresses on an unyielding surface be
neath a pneumatic tire (U.S. Army Waterways Experiment Station Report, Vick
sburg, Miss.).
[43] Hofelt, C., Factors that influence skid resistance. Part V: Effect of speed, load distribu
tion, and inflation, Proc. First International Skid Prevention Conference, Char-
lottesville, Va., Aug. 1959.
[44] Clark, S. K., "An Experimental Study of Pressure Distribution Curves Applicable to
Pneumatic Tires", NASA Contractors Report CR-754, National Aeronautics and
Space Administration, Washington, D. C., May 1967.
[45] Robbins, D. H., The contact of certain elastic shells with rigid flat surfaces. Report
05608-7-T, (Office of Research Administration, The University of Michigan, Ann
Arbor, Mich., Sept. 1965).
[46] Akasaka, T., On the bending of a tire, Series B-9, Report No. 19, Chuo University, To
kyo, Japan.
[47] Hertz, H. J.. Reine angew. Math. 92, 156 (1886). See also Timoshenko, S. P., and God-
dier, J. N., Theory of Elasticity (McGraw-Hill Book Co., New York, 1956).
[48] Frank, F., and Hofferberth, W., Mechanics of the pneumatic tire, Rubber C hem. Tech.
40(1), 271 (Feb. 1967).
[49] Martin, F., Theoretische Untersuchungen zur Frage des Spannungszustandes in Luf-
treifen bei Abplattung, Jahrbuch der deutschen Luftfahrforschung, Teil I, 470
(1939).
[50] Vlassov, V. Z., Izvest. Akad. Tekh. Nauk. No. 6, 819 (1949).
[51] Böhm, F., Zur Mechanik des Luftreifens, Habilitationschrift th., Stuttgart, 1965.
[52] Zakaharov, S. P., and Novopol'skii, Distribution of specific pressure of a tire on the
road at high velocities (in Russian), Trudy Tauchno-Issledovatel'skoga Instituta
Shinnoi Pronyschlennosti, Sbornik 3, Methody Rascheta i Ispaytaniya Autorao-
bil'nykh Shin: U.S.S.R. Nauchno, Moscow: 1957, p. 139-53; Rubber Abs. 36, 439
(Sept. 1958).
[53] Essenburg, F., and Gulati, S. T., On the contact of two axisymmetric plates, J. Appl.
Mech. 33 (2), 341 (June 1966).
[54] Novopol'skii, V. I., and Nepomnyashchii, E. F., The interaction of a motor vehicle
tyre tread with the road surface, Abrasion of Rubber, D. I. James, ed., (Palmerton
Publishing Co., New York).
[55] Biderman, V. L., Volodina. T. N., and Pugin, V. A. Determining the relationship be
CONTACT BETWEEN THE TIRE AND ROADWAY 359

tween the deformation of a tyre carcass and the displacement of the carcass relative
to the road, Sov. Rubber Tech. 25 (7), 37 (July 1966).
[56] Sen/. N. and Hussman, A. W., "Forces and Displacement in Contact Area of Free
Rolling Tires", Soc. Auto. Engr's Paper 710625, June 1971.
[57] Gough, V. E., Tyre to ground contact stresses. Wear 2, 107 (1958/59).
[58] Gough, V. E., Practical tyre research, SAE Trans. 64, 310 (1956).
[59] Cooper, D. H . Distribution of side force and side slip in the contact area of the pneu
matic tire, Kautschuk and Gummi 11 (10), WT 273-277 (Oct. 1958).
[60] Iritani, S., and Baba, T., Forces on the contact patch of the tyre, Paper No. 3.1, Tenth
International FISITA Congress, Tokyo, Japan, 1964.
[61] Vermeulen, P. J., and Johnson, K. L... Contact of nonspherical elastic bodies trans
mitting tangential forces, J. Appl. Mech. 31 (2), 338 (June 1964).
[62] Saito, Y., A study of dynamic steering properties of pneumatic tires, Proc. 10th FI
SITA Congress, Tokyo, 1964.
[63] Mindlin, R. D., Mechanics of Granular Media, Proc. Second U.S. Nat. Cong. Appl.
Mech. (1954). Published by A.S.M.E., New York.
[64] Novopol'skii, V. I., and Tret'yakov, O. B., Slip of the elements of the tread pattern in
the contact area of tires, Sov. Rubber Tech. 22 (11), 25 (Nov. 1963).
[65] Home. W. B., Yager, T. J., and Taylor, G. R., Recent research on ways to improve tire
traction on water, slush, or ice. Paper presented at AIAA Aircraft Design and Tech
nology Meeting, Los Angeles, Calif., Nov. 15-18, 1965.
[66] Kelley, J. D., and Allbert, B. J., Tread design of tire affects wet traction most, SAE
Journal 76 (9), 58 (Sept. 1968).
[67] Bowden, F., and Tabor, D., Friction and Lubrication of Solids, Vol. 2, (Oxford Univ.
Press, New York, 1964).
68 Ref. [3]. p.
69' Maycock, G., Proc. IME 180-2A (4), 122 (1956-66).
70 Daube, J., La Technique Routiere 4 (1), 4 (1959).
71 Leland, T., and Taylor, R., J. Aircraft 2 (2), 72 (1965).
72 Grosch, K., Proc. Roy. Soc. A274, 21 (1963).
73 Ludema, K., and Tabor, D., Wear 9, 329 (1966).
M Schallamach, A., Wear 7, 375 (1963).
la Savkoor, A., Wear 9, 66 (1966).
76] Bartenev, G., and Elkin, A., Wear 8, 8 (1965).
77 Krempel, G., A. T. Z. 69 (1), 128 (Jan. 1967, A. T. Z. 69 (8), 1334 (Aug. 1967).
78' Hurry, J. A., and Prock, J. D., India Rubber World 128, 619 (1953).
79' Atach, D., and Tabor, D., Proc. Roy. Soc. 236, 539 (1958).
80' Ludema, K., and Lee, C. S., ASME Paper No. 69-LUB-20 (1969).
81 Bevilacqua, E., and Percarpio, E., Rubber Chem. Tech. 41 (4), 832-894 (Sept. 1968).
82 Holland, L., Properties of Glass Surfaces (John Wiley & Sons, Inc., New York, 1964).
83] Cameron, A., Principles of Lubrication (John Wiley & Sons, Inc., New York, 1966).
84] Moore, D., Wear 13, 381 (1969).
85] Roberts, A., and Tabor, D., Nature 219, 1122 (1968).
86] Leland, T. J. W., Proc. 9th Stapp Car Crash Conf., p. 383, University of Minnesota,
1966.
[87] Nybakken, G., Staples, R., and Clark, S., Laboratory Experiments on reverted rubber
friction, Tech. Report 7 (Office of Research Administration, The University of
Michigan, Ann Arbor, Mich., Oct. 1968).
88] Moore, D. F., Research Trends, p. 8 (Spring-Summer 1966).
89] Moore, D. F., Wear 8, 245 (1965).
90] Wehner, B., Her Bauingenieur 10 (I) (1965).
91] Schulze, K. H., Strasse und Autobahn 10 (10), 379 (1959).
92] Schallamach, A., Rubber Chem. Tech. 27, 439 (1954).
93] Goodenow, G., Kolhoff, T., and Smithson, F., SAE Paper No. 680137, 13 (Jan. 1968).
94] Holmes, K. E., and Stone, R. D., Symposium on Handling of Vehicles Under Emer
gency Conditions, Inst. Mech. Engrs , Jan. 8, 1969, p. 81.
[95] Dugoff, H., and Brown, B. J., Instrumentation for measuring tire shear forces, SAE Pa
per No. 700093 (Jan., 1970).
[96] Browne, A. L., "Mathematical Analysis for Pneumatic Tire Hydroplaning," Surface
Texture Versus Skidding: Measurements, Frictional Aspects, and Safety Features of
Tire-Pavement Interactions, ASTM STP 583, American Society for Testing and Ma
terials, 1975, pp. 75-94.
[97] Holmes, T., Lees, G., and Williams, A. R., "A Combined Approach to the Opti
mization of Tyre and Pavement Interaction," Wear, Vol. 20, 1972, pp. 241-276.
[98] Ludema, K. C. and Gujrati, B. D., An Analysis of the Literature on Tire-Road Skid
Resitance, American Society for Testing and Materials Special Technical Pub
lication 541, Philadelphia, 1973.
360 MECHANICS OF PNEUMATIC TIRES

[99] Sinnamon, J. F., "Literature Survey of Tire-Road Experiments," Highway Safety Re


search Institute—University of Michigan Report No. UM-HSRI-PF-74-5, Ann Ar
bor, 1974.
[ 100] Holla, L. and Yandell, W. O., "A Review of Research on Road-Tyre Friction," Aus
tralian Road Research, Vol. 5, No. 2, June 1973, pp. 76-91.
[101] Meyer, W. E. and Schrock, M. D., 'Tire Friction, a State of the Art Review," The
Pennsylvania State University Automotive Safety Research Program Report S-34,
April 1969.
| U)2] Giles, C. G., Highway Research Record. No. 46, p. 43, 1964.
[103] van Eldik Thieme, H. C. A., "Cornering and Camber Experiments," Section 7.3 of
NBS Monograph 122, Mechanics of Pneumatic Tires, ed. by S. Clark, p. 660, Nov.
1971.
[104] Home, W. B., Yager, T. J., and Taylor, G. R., "Review of Causes and Alleviation of
Low Tire Traction on Wet Runways," NASA TN D-4406, April 1968.
[105] Li, W. H. and Lam, S. H., Principles of Fluid Mechanics, Addison-Wesley, Reading,
Massachusetts 1964.
[106] Allhert. B. J., 'Tires and Hydroplaning," SAE Paper No. 680140, 1968.
[107] Home, W. B., Yager, T. J. and Taylor, G. R., "Recent Research on Ways to Improve
Tire Traction on Water, Slush or Ice," AIAA Paper No. 65-749, presented at the
AIAA Aircraft Design and Technology Meeting, November, 1965.
[108] Nybakken, G. H., Staples, R. J., and Clark, S. K. , "Laboratory Experiments on Re
verted Rubber Friction," NASA Cr-1398, August, 1969.
[109] Schallamach, A., "Skid Resistance and Directional Control" Chapter VI of NBS
Monograph 122, Mechanics of Pneumatic Tires, edited by S. K. Clark, Washington
1971, pp. 538-539.
[110] Gough, V. E., Hardman, J. H., and Mac Laren, R. J., "Abraded Tire Treads," Inst.
Rubber Ind. Trans. Vol. 32, 1956, pp. 1-27.
[111] Stock er, A. J. and Lewis, J. M., "Variables Associated with Automobile Tire Hydro
planing," Texas Transportation Institute-Texas A & M University Research Report
147-2 Study TTI-2-8-70- 147-2, Sept. 1972.
[112] Gengenbach, W., "The Effect of Wet Pavement on the Performance of Automobile
Tires," Ph.D. thesis, University of Karlsruhe, July 1967.
[113] Sinnamon, J. F. and Tielking, J. T., "Hydroplaning and Tread Pattern Hydro
dynamics," Highway Safety Research Institute-University of Michigan, Report
UM-HSRI-PF-74-10.
[114] Meades, J. K., "Braking Force Coefficients Obtained with a Sample of Currently
Available Radial Ply and Crossed Ply Car Tires," Road Research Laboratory Re
port LR-73, 1967.
[1 15] Staughton, G. C., "The Effect of Tread Pattern Depth on Skidding Resistance," Road
Research Laboratory Report LR-323, 1970.
[116] Staughton, G. C. and Williams, T., "Tyre Performance in Wet Surface Conditions,"
Road Research Laboratory Report LR-355, 1970.
[117] Moore, D. F., The Friction of Pneumatic Tires, Elsevier, New York, 1975.
[118] Bond, R., Lees, G. and Williams, A. R., "An Approach Towards the Understanding
and Design of the Pavements Textural Characteristics Required for Optimum Per
formance of the Tyre," in The Physics of Tire Traction, Theory and Experiment, D.
F. Hays and A. L. Browne, Eds., Plenum Press, New York, 1974 pp. 325-338.
[1 19] Smith, R. W., Rice, J. M., and Spelman, S. R., "Design of Open-Graded Asphalt Fric
tion Courses," Federal Highway Administration Report No. FHWA-RD-74-2,
Washington, January, 1974.
[120] Mullen, W. G. and Dahir, S. H. M., "Skid Resistance and Wear Properties of Aggre
gates for Paving Mixtures," North Carolina State University Highway Research
Program Project ERD-1 10-69-1, Interim Report, June 1970.
[121] Kummer, H. W., "The Polishing of Road Stones on a Drum Apparatus," The Pennsyl
vania State University, Automotive Safety Research Program, Report No. 317, Uni
versity Park, Pennsylvania, April 1968.
[122] I'Anson, R., "An Investigation of Dynamic Aquaplaning Using Small Pneumatic
tyres," Ph. D. thesis. University of Bristol, 1973.
[123] Hegmon, R. R., Werner, S., and Runt, L. J., "Pavement Friction Test Tire Correla
tion," Federal Highway Administration Report No. FHWA-RD-75-88, Washing
ton, April 1975.
[124] Nordstrom, O., "The Effect of the Depth of Water Film on Road Friction," Chapter 3,
Prague Report of the Technical Committee on Slipperiness, Nov. 1970.
[125] Leland, T. J. W., "An Evaluation of Some Unbraked Tire Cornering Force Character
istics," NASA Technical Note NASA Tn D-6964, Washington D. C., Nov. 1972.
[126] Gallaway, R. M., Schiller, R. E. Jr., and Rose, J. G., "The Effects of Rainfall Intensity,
CONTACT BETWEEN THE TIRE AND ROADWAY 361

Pavement Cross Slope, Surface Texture, and Drainage Length on Pavement Water
Depth," Texas Transportation Institute, Texas A & M University Research Report
No. 138-5, 1972.
[127] Dunlop, D. F., Fancher, P. S., Scott, R. E. MacAdam, C. C., and Segal, L., "Influence
of Combined Highway Grade and Horizontal Alignment on Skidding," Highway
Safety Research Institute, University of Michigan, Report on NCHRP Project 1-14,
1975.
[128] Preus, C. K., "Effects of Studded Tires on Pavements and Traffic Safety in Minne
sota," Society of Automotive Engineers, Paper No. 720117, Jan. 1972.
[129] Saal, R. N. J., "Laboratory Investigations into the Slipperiness of Roads," Chemistry
and Industry, Vol. 55, Jan. 1936, pp. 3-7.
[130] Hays, D. F., "Squeeze Films for Rectangular Plates," ASME Paper 62- S-9, American
Society of Mechanical Engineers, New York, 1962.
[131] Moore, D. F., "On the Inclined Non-inertial Sinkage of a Flat Plate," J. Fluid Mech.,
20 (1964), pp. 321-330.
[132] Moore, D. F., "The Sinkage of Flat Plates on Smooth and Rough Surfaces," Doctoral
Dissertation, Mechanical Engineering Department, Pennsylvania State University,
1963.
[133] Mosher, L., "Results from Studies of Highway Grooving and Texturing by Several
State Highway Departments," Paper No. 27 of the Conference on Pavement Groov
ing and Traction Studies, Langley Research Center, Hampton, Virginia, Nov. 18-
19, 1968; published in NASA Sp-5073.
[134] Dawson, D. and Higginson, G. R., Elastohydrodynamic Lubrication, Pergamon Press,
New York, 1966.
[135] Browne, A., Cheng, H., and Kistler, A., "Dynamic Hydroplaning of Pneumatic Tires,"
Wear, Vol. 20, pp. 1-28, 1972.
[136] Browne, A. L., "Predicting the Effect of Tire Tread Design on Thick Film Wet Trac
tion," Tire Science and Technology, TSTCA, Vol. 5, No. 1, February 1977, pp. 6-
28.
[137] Bathelt, H., "Calculation of the Aquaplaning Behavior of Smooth and Profiled tires,"
Automobil Technische Zeitschrift International, Vol 75, No. 10, 1973, pp. 12-17.
[138] Whicker, D., Browne, A. L., and Rohde, S. M., "Some Effects of Inclination on Elas
tohydrodynamic Squeeze Film Problems," Journal of Fluid Mechanics, Vol. 78,
Nov. 23, 1976, pp. 247-260.
[139] Gallaway, R. M., Rose, J. G., Scott, W. W. Jr., and Schiller, R. E., "Influence of Water
Depths on Friction Properties of Various Pavement Types," Texas A & M—Texas
Transportation Institute Research Report 138-6, August 1974.
[140] Veith, A. G. and Pottinger, M. G., 'Tire Wet Traction: Operational Severity and its
Influence on Performance, in The Physics of Tire Traction, Theory and Experiment,
D. F. Hays and A. L. Browne, Eds., Plenum Press, New York, 1974, pp. 5-24.
[141] Meades, J. K., "The Effect of Tyre Construction on Braking Force Coefficients," Road
Research Laboratory Report LR 224, Crawthorne, Berkshire, 1969.
[142] Peterson, R. F. Jr., Eckert, C. F., and Carr, C. E., "Tread Compound Effects of Tire
Traction," in The Physics of Tire Traction, Theory and Experiment, D. F. Hays and
A. L. Browne, Eds., Plenum Press, New York, 1974, pp. 223-239.
[143] Gough, V. E., "A Tyre Engineer Looks Critically at Current Traction Physics," in The
Physics of Tire Traction, Theory and Experiment. D. F. Hays and A. L. Browne, Eds.,
Plenum Press, New York 1974, pp. 281-297.
[144] Rose, J. G., Hutchinson, J. W., and Gallaway, B. M., "Summary and Analysis of the
Attributes of Methods of Surface Texture Measurement," in Skid Resistance of
Highway Pavements, ASTM STP 730, American Society for Testing and Materials,
1973, pp. 60-77.
[145] Doty, R. N., "Study of the Sand Patch and Outflow Meter Methods of Pavement Sur
face Texture Measurement," in Surface Texture Versus Skidding: Measurements,
Frictional Aspects and Safety Features of Tire-Pavement Interactions, ASTM STP
583, American Society for Testing and Materials, 1975, pp. 42-61.
[146] Henry, J. J. and Hegmon, R. R., "Pavement Texture Measurement and Evaluation,"
Surface Texture Versus Skidding: Measurements, Frictional Aspects, and Safety Fea
tures of Tire-Pavement Interactions, ASTM STP 583, American Society for Testing
and Materials, 1975, pp. 3-17.
[147] Schonfeld, R., "Pavement Surface Texture Classification and Skid Resistance Photo—
Interpretation," in The Physics of Tire Traction: Theory and Experiment, D. F. Hays
and A. L. Browne, Eds., Plenum Press, New York, 1974, pp. 325-338.
[148] Veres, R. E., Henry, J. J., and Lawther, J. M., "Use of Tire Noise as a Measure of
Pavement Macrotexture," Surface Texture Versus Skidding: Measurements, Fric
tional Aspects, and Safety Features of Tire-Pavement Interactions, ASTM STP 583,
American Society for Testing and Materials, 1975, pp. 18-28.
362 MECHANICS OF PNEUMATIC TIRES

[149] Grime, G. and Giles, C. G., "The Skid Resisting Properties of Roads and Tires," Pro
ceedings of the Automobile Division, Institution of Mechanical Engineers, 1954-55,
No. 1, p. 19-39.
[150] , "Delft Tire Tread Drainage Meter," Leaflet Number 5, Vehicle Research Lab
oratory Delft University of Technology 1972.
[151] Tabor, D., "Friction Between Tyre and Road," Engineering, Dec. 26, 1958, pp. 838-
840.
[152] Yandel, W. O., "The Relation Between the Stress Saturation of Sliding Rubber and
the Load Dependence of Road Tyre Friction" in The Physics of Tire Traction, The
ory and Experiment, D. F. Hays and A. L. Browne, Eds., Plenum Press, New York,
1974, pp. 311-323.
[153] Barnes, P., D. Tabor and J. C. F. Walker, "The Friction and Creep of Polycrystalline
Ice," Proc. Roy. Soc., London A, 324 p. 127-155 (1971).
154] Southern, E. and R. W. Walker, "Nature" Physical Series, v. 237, p. 142 (1972).
155] Gnovich, W. and K. A. Grosch, "Journal of IRI," p. 192, 1972.
156] Kalker, J. J. and A. D. de Pater, Eds., "The Mechanics of the Contact Between Defor-
mable Bodies," p. 1-25, Delft University Press, 1975.
[157] Mover, R. A., "Braking and Traction Tests on Ice, Snow, and on Bare Pavements,"
Highway Research Board, Proceedings of the 27th Annual Meeting, Washington D.
C, Dec. 2-5, 1947, pp. 340-360.
[158] Anon., Snow Removal and Ice Control Research, Highway Research Board Special Re
port 115, 1970.
[159] Browne, A. L., "Tire Traction on Snow-Covered Pavements," in The Physics of Tire
Traction, Theory and Experiment," D. F. Hays and A. L. Browne, Eds., Plenum
Press, New York, 1974, pp. 99-139.
[160] Heumann, G., "Germany Bans Studded Snow ires," Machine Design, Feb. 6, 1975,
pp. 30-33.
[161] Ichihara, K. and Mizoguchi, M., "Skid Resistance of Snow- or Ice-Covered Roads," in
Snow Removal and Ice Control Research, Highway Research Board Special Report
115, 1970, pp. 104-114.
[162] Smith, R. W. and dough, D. J., "Effectiveness of Tires Under Winter Driving Condi
tions," presented at the 51st Annual Meeting of the Highway Research Board, Jan.
1972.
163 Grosch, K. A. and Maycock, G., Trans. I. R. I., Vol. 42, T 280, 1966.
164 Anon., "Safer Winter tires," Rubber Developments, Vol. 20, No. 3, pp. 82-85, 1967.
165 Frere, P., "Better Than Studded Tires," Motor, pp. 39 week ending January 19, 1974.
166] Setright, L. J. K., "How Long is a High-Speed tyre," Automotive Engineer, Vol. 1, No.
4, pp. 15-17, 1976.
[167] Sapp, T., "Ice and Snow tire Traction," Society of Automotive Engineers, Paper No.
680139, Jan. 1968.
[168] Reinhart, M. A., "Factors in Tires that Influence Skid Resistance; Part IV: The Effect
of Tread Composition," in First International Skid Prevention Conference, Pro
ceedings, Part I, pp. 167-173, Aug. 1959.
[169] Anon., "Snow tires," Consumer Bulletin, Vol. 41, No. 12, pp. 39, 27-29, Dec. 1958.
[170] Tomarkin, L. W., "Increasing Tire Traction on Wet, Snowy and Icy Surfaces; Tomar-
kin Process," Rubber Age, v. 83, pp. 832-835, August 1958.
[171] Schaerer, P. A., "Compaction or Removal of Wet Snow by Traffic," in Snow Removal
and Ice Control Research, Highway Research Board Special Report 115, 1970, pp.
97-103.
[172] Anon., "1957 Winter Test Program; Committee on Winter Driving Hazards," Traffic
Safety Research Review, v. 2, No. 3, Sept. 1958, pp. 26-31.
[173] Automobil Club der Schweiz, Berne, Winterreifen Prufung (1962).
[174] Anon., "Superhard Particles Embedded in Tire Rubber Said to Increase Traction,
Stopping Ability," Highway Research News, No. 47, Spring 1972, pp. 37-39.
[175] Hegmon, R. R. and Meyer, W. E., "The Effectiveness of Antiskid Materials," Highway
Research Record Number 227, 1968, pp. 50-56.
[176] Johnson, W. C., "The Effect of Carcass Construction, Size, Cord Angle, and Number
of Plies," Proceedings of the first International Skid Prevention Conference, Virginia
Council of Highway Investigation and Research, Charlottesville, 1959, Part I, pp.
163-166.
[177] Schuring, D. J. and Roland, R. D., "Radial Ply Tires—How Different are They in the
Low Lateral Acceleration Regime," SAE Paper No. 750404, Feb. 1975.
[178] Maycock, G., "Experiments on Tire Tread Patterns," Road Research Laboratory Re
port LR 122, 1967.
[179] Kelley, J. D., Jr., "Factors Affecting Passenger Tire Traction on the Wet Road," SAE
Paper No. 680138, 1968.
CONTACT BETWEEN THE TIRE AND ROADWAY 363

[180] Neill, A. H. Jr., "Wet Traction of Tractionized Tires," National Bureau of Standards,
Technical Note 566, 1971.
[181] Campbell, K. L., Spelman, R. H., Tarpinian, H. D., and Johnson, D. E., "SAE Tire
Committee Studies Effect of Tire Wear on Wet Pavement Traction," SAE Journal,
July 1970, pp. 44-47.
[182] Dijks, A., "A Multifactor Examination of Wet Skid Resistance of Tires," SAE Paper
No. 741106.
[183] Dijks, A., "Tests on the Minimum Permissible Tread Depth of Passenger Car Tires,"
ATZ, Vol. 75, No. 1, 1973, pp. 1-6.
[184] Kern, W. F., "Coefficient of Wet Friction of Tire Treads," Rubber Chemistry and
Technology, Vol. 40, 1967, pp. 984-1013.
[185] Holmes. K F... "Braking Force/Braking Slip Measurements Over a Range of Condi
tions Between 0 and 100 Percent Slip," Road Research Laboratory Report LR 292,
1970.
[186] Grosch, K. A. and Maycock, G., "Influence of Test Conditions on Wet Skid Resis
tance of Tire Tread Compounds," Rubber Chemistry and Technology, Vol. 41,
1968.
[187] Farber, E. et. al., "Determining Pavement Skid—Resistance Requirements at Inter
sections and Braking-Sites," NCHRP Report No. 154, Franklin Institute, Phila
delphia, 1974.
[188] Perkins, T. H. and Bagi, D. J . "An Investigation of the Relationship Between Traffic
Volume and Accident Frequency at Rural Intersections," Bureau of Traffic, Divi
sion of Highways, Ohio Department of Transportation, Columbus, Ohio, 1974.
[189] , "A Policy on Geometric Design of Rural Highways, 1965," American Associa
tion for State Transportation Officials, Washington, D. C, 1966.
[190] , "A Policy on Design of Urban Highways and Arterial Streets, 1973," Ameri
can Association for State Transportation Officials, Washington, D. C., 1974.
[191] Bekker, M. G. and Semonin, E. V., "Motion Resistance of Pneumatic Tyres," Journal
of Automotive Engineering, April 1975, pp. 6-10.
[192] Bekker, M. G., Theory of Land Locomotion, The University of Michigan Press, Ann
Arbor, 1956.
[193] Bekker, M. G., Off-the-Road Locomotion, The University of Michigan Press, Ann Ar
bor, 1960.
[194] Bekker, M. G., Introduction to Terrain-Vehicle Systems, The University of Michigan
Press, Ann Arbor, 1969.
[195] Czako, T. F., "Methods of Vehicle Soft Soil Mobility Evaluations Based on the Soil
Vehicle Interaction," in Proceedings of the First International Conference on Vehicle
Mechanics, H. K. Sachs, Ed., Wayne State, Detroit, 1969.
[196] Janosi, /., "Theoretical Analysis of the Performance of Tracks and Wheels Operating
on Deformable Soils," Transactions of the ASME, 1962, pp. 133-134.
[197] Satake, M. and Mukai, T., "Traction and Flotation Characteristics of liarth mover
Tires on Soft Soil," SAE Paper No. 720743, 1972.
[198] Pules, M. L. and Eves, D. J., "ATV Flotation Tires," SAE Paper No. 720765, 1972.
[199] Semonin, E. V. and Wilson, M. A., "Twenty-Five Years of Progress in Tires for Earth-
moving Equipment," SAE Paper No. 740416, 1974.
[200] Vermie, H. R., "The ABC's of Radial Off-the-Road Earthmover Tires," SAE Paper
No. 740679, 1974.
[201] Trindal, W. S., "Technical Analysis Study of Off-Road Tires," SAE Paper No. 730853,
1973.
[202] TRA Yearbook and Supplementary Service Data Book, The Tire and Kim Association,
Inc., Akron, Ohio.
[203] MacFarland, R. M., "Recommended Practice for Tire TMPH Application," SAE Pa
per No. 730855, 1973.
Chapter 6
TIRE TRACTION AND WEAR
A. Schallanmach1
K. Grosch2

6. 1. Rubber Friction 367


6.1.1 Load dependence 367
6.1.2 The effect of sample shape 369
6. 1.3 Hard sliders on rubber and the frictional lift 370
6.1.4 The temperature and speed dependence of rubber
friction 375
6.1.5 The sliding mechanism 380
6. 1 .6 The surface condition of sliding rubber 385
6. 1 .7 Rubber friction on lubricated surfaces 388
6.1.8 Hydrodynamic lubrication 391
6. 1 .9 The friction coefficient on ice 394
6.1.10 Filler effects and oil extension 396
6. 1 . 1 1 The effect of polymer blending 398
6.1.12 The range of the aTv values in tire skids 400
6.2. Abrasion 402
6.2.1 Introduction 402
6.2.2 The abrasion pattern 403
6.2.3 The load dependence 406
6.2.4 Temperature and speed dependence 407
6.2.5 The effects of ant ioxidants and the surrounding atmo
sphere 411
6.2.6 Abrasion by a blade 414
6.3. Tire forces on dry and wet surfaces 415
6.3.1 Introduction 415
6.3.2 Simple slip 416
' Formerly with The Malaysian Rubber Producers Research Association, Brickondonbury, Herts., England
Now retired. Present address 74 Longmore Avc . Barnet, Herts., England.
2 Formerly with Uniroyal A.G. Aachen, Germany
Now with Continental Gummiwerk A.G., Aachen, Germany.

365
366 MECHANICS OF PNEUMATIC TIRES

6.3.3 Load—slip equivalence.... 421


6.3.4 Composite slip 422
6.3.5 Cornering 423
6.3.6 The eflect of load transfer 427
6.3.7 Side force and self-aligning torque at varying slip
angle 428
6.3.8 The influence of road conditions on side force and self
aligning torque 431
6.3.9 The influence of tire construction on side force and
self-aligning torque 439
6.4. Braking and traction of tires 441
6.4.1 Experimental methods 441
6.4.2 The skid coefficient vs. time curve 442
6.4.3 The slip dependence of the skid coefficient 443
6.4.4 The speed dependence of the peak and sliding value
of the skid resistance 444
6.4.5 Tread pattern and surface roughness effects 447
6.4.6 The effect of siping 449
6.4.7 The effect of tire construction 449
6.4.8 The compound eflect 451
6.5. Tire Wear 454
6.5.1 Theory of tire wear 454
6.5.2 Load and slip dependence of wear 457
6.5.3 Eflect of tire construction 459
6.5.4 The relative wear rating of tread compounds 461
6.5.5 Uneven tire wear 467
TIRE TRACTION AND WEAR 367

6.1. Rubber Friction


The laws of rubber friction differ from those of solid friction in that the
coefficient of friction depends on normal load, sliding speed and temper
ature. All these effects are also influenced by the surface topology of the
track. Speed and temperature dependence are, however, intimately con
nected; their mutual relation is both theoretically and practically the most
important result of research into rubber friction.
6.1.1 Load dependence
The friction coefficient decreases with increasing load, most pro
nouncedly so on smooth tracks [1, 2]. Figure 6.1 shows results for three
unfilled NR compounds of different stiffness on plate glass; similar data
have been obtained on smooth ice [3]. The large friction coefficients evi
dent from figure 6. 1 are produced without measurable abrasion so that the
frictional bonds between rubber and track must be reversible, in funda
mental contrast to solid friction which is a concomitant of abrasion; but, as
in solid friction, the load dependence of rubber friction can be explained
with the assumption that the friction force F is proportional to the true
contact area A which is determined by the asperities on both friction part
ners and by their deformation under load [4, 2]
F = <pA (6.1)
where <p is a material constant depending on speed and temperature. With
an ideally smooth track, only the rubber asperities need be considered. If
these asperities are taken to be closely spaced and identical, dimensional
reasoning leads to the following expression for the friction coefficient /i [5]

c.(p/Efn (6.2)

where p is the normal pressure and E is the Young's modulus of the rub
ber, cn and /?„ are constants. The essential prediction of eq. (6.2) is that the
quantity (£/x/<p) should be a universal function of the ratio (p/E). Eq. (6.2)
can be made more specific by ascribing spherical shape to the asperities.
Using Hertz's equations for contact pressures, the friction coefficient be
comes
M = (<?/E)(p/E)-»\ (6.3)
The full curves in figure 6. 1 show how far eq. (6.3) conforms to experi
ment.
At high pressures, the asperities bulge out, interfere with each other and
eq. (6.3)—but not eq. (6.2)—loses validity. In order to follow this process,
a square array of 25 small rubber hemispheres was compressed by a glass
plate, as shown in figure 6.2, and their contact area was measured [5]. Un
der high loads, the asperities are squeezed into a solid block, and the con
tact area hardly changes on further increasing the load. The friction coef
ficient should then become inversely proportional to the load, according to
eq. (6.1). The solid line in figure 6.3 gives the theoretical friction coeffi
cient as derived from the model experiment in figure 6.2 and from eq.
(6.1). The experimental points were plotted as Eμ = f (p/E) and superim-
368 MECHANICS OF PNEUMATIC TIRES

Modulus E
8.9 kgf/cm2
• 18.3
026.9

p, kp/cm

FIGURE 6. 1. Pressure dependence of the friction coefficient of unfilled NR with the indicated
Young's moduli E on polished glass V = 0.0021 cm/s; Thefull curvesfollow the equation fi —
const. p~'/3 (from ref. PJ).

posed on the theoretical curve, the vertical displacement giving the factor
<i>. Experiment and theory agree, apart from pressures below about 0.04 E
where the points fall on a flatter curve. This divergence is most probably
due to the asperities being statistically distributed in size, as detailed in
reference [5]. Any deviation from a uniform surface topology reduces the
pressure dependence of rubber friction.
Friction on rough tracks can be similarly approached. If the track asper
ities are assumed to be hemispheres of much larger size than the rubber
asperities, the friction coefficient becomes [5]
const. (<p/E)(p/E)- (6.4)
where the constant contains the height-to-width ratio of the track asper
ities. The friction coefficient on rough tracks should thus be much less
pressure dependent than on smooth tracks. The results in figure 6.4 for
friction on silicon carbide paper follow eq. (6.4) until high pressures are
reached when /i decreases faster than predicted. This could be caused by
saturation effects like those in figure 6.2, and by rubber filling the gaps be
tween track asperities.
Eq. (6.4) does not contain the asperity size so that ju should be independ
ent of track roughness as long as the grains are geometrically similar [6].
Table 6.1 shows that μ decreases at most by about 15% for a nearly 5-fold
increase in grain size, and that it depends little on load. It will therefore be
TIRE TRACTION AND WEAR 369

^V ^^f ^^ ^^ '^^T

• ••••
I

FIGURE 6.2. Contact area of a model rubber surface under the pressure of: (a) 2.14;(b) 7.75;
(c) 66.0 kgf/cm2 (from ref. [5]).

sufficient for many purposes to consider the coefficient of rubber friction


as constant unless pressure concentrations occur in the contact (see next
section).
The near agreement between theory and experiment in figure 6.4 does
not prove that friction on rough tracks is, like friction on smooth tracks,
solely determined by the true contact area. If it were so, temperature and
velocity dependence on both types of track should be identical, contrary to
the experimental evidence presented in section 6.1.4.

6.1.2 The effect of sample shape [7]


Friction experiments are mostly made with squat samples which are tac
itly assumed to undergo pure shear, as sketched in figure 6.5 (a). Pure
shear would, however, need the distributed forces F3 and F4 on the free
surfaces. In their absence, the forces /-", and /•'.. form a couple tilting the
test piece forward; the front edge is pulled into the contact, and the rear
edge is lifted off the track, as indicated in figure 6.5 (b). The deformation
resembles the bending of a short cantilever. The load is now carried by the

-0.5
-2.0 -1.5 -1.0 -0.5
LOG|0(p/E)

FIGURE 6.3. Theoretical and experimental pressure dependence of the coefficient offriction
on smooth tracks (from ref. [5]). Points A, B, C from (3); Efrom (I) unknown modulus,
adjustedfor best fit.
370 MECHANICS OF PNEUMATIC TIRES

(a)
02
01
0
-O.I (b)
=L-Q2
o 0.1

3 °
-O.I
Q2
at
-0.4 -0.2 0 0.2 0.4 0.6 0.8 IX)
LOG,0p

FIGURE 6.4. Pressure dependence of the friction coefficient of rubber compounds on silicon
carbide paper at 0.0021 cm/sec. (a) unfilled natural rubber; (b) natural rubber tire tread; (c)
styrene-butadiene tire tread (from ref. (5J).

front part where the normal pressure is raised well above its nominal
value. The intensity of this effect increases with increasing height-to-
length ratio of the test piece. Figure 6.6 shows the contact of square sam
ples of different height sliding on Plexiglass; the increased pressure on the
remaining contact reduces the friction coefficient, as expected from the
preceding section.
Samples trisected by two parallel cuts ("sipes"), shown in figure 6.7, in
troduce a new phenomenon. When sliding in the direction of the cuts, ad
ditional contact is established near the center, and friction is greater than
with an uncut sample. This has been ascribed to buckling. A compressive
stress develops parallel to the cuts because friction concentrated near the
front tends to hold the test piece back while its rear part is brought for
ward by the imposed sliding velocity. By analogy with the instability crite
rion of loaded columns, the probability of buckling will increase with de
creasing width of the rubber between the cuts; ie. with increasing number
of cuts. The friction coefficient of samples divided into 5 parts rose to 1.44.

6.1.3 Hard sliders on rubber and the frictional lift


The surface of rubber sliding on coarse tracks is deformed both nor
mally (indentation) and tangentially; the work done in maintaining the
corresponding strain rate is part of the measured friction. The tangential

TABLE 6. 1 Coefficients offriction on garnet paper ofdifferent coarseness at a sliding speed of


0.35 cm/sec. (from [6])

Natural rubber Styrene-butadiene


Compound tire tread tire tread
Pressure, kgf/cm2 0.53 1.84 0.53 1.84
Grain size, mm 0.13 1.42 1.40 1.41 1.41
0.29 1.30 1.31 1.39 1.40
0.63 1.21 1.27 1.29 1.36
TIRE TRACTION AND WEAR 371

FIGURE 6.5. Deformation ofsliding rubber blocks; (a) due to shear only, and (b) due to shear
and bending. From ref. ff] and pi].

component, which is the more important component on dry tracks, de


pends on the steepness of the leading flank of the asperities. Model experi
ments were made with hard sliders on rubber -to throw light on the effect
of grain shape on friction [8]. To avoid difficulties with abrasion, the sli
ders were cylindrical or prismatic; their cross sections are shown in figure
6.8. The chisel (b) could be slid in either direction, producing angles of at
tack (= angle between leading flank and rubber) of 25° and 90°. Figure
6.9 gives results obtained on two different tracks, expressed as friction ra
tios (= tractive force/normal force) to distinguish them from the friction
coefficient /i between flat surfaces. On elementary reasoning ju should be
come infinite when the cotangent of the angle of attack is equal to, or
smaller than /x; in fact, an angle of 90° gives only about twice the friction
at 25°.
The friction ratio is kept finite by the extension of the rubber behind the
slider which helps to pull it around its edge. This strain is demonstrated in
figure 6.10 showing the distortion by a spherical slider of a rubber surface
with a square lattice marked on it [9]. The rubber is extended behind, and
compressed in front of the slider. Friction at the edge of a prismatic slider
resembles that of a rope wound around a pulley or capstan which in
creases exponentially with the wrapping angle.
In an elementary theoretical treatment, the edge of the slider was as
sumed to be cylindrical, and the friction coefficient to be constant. The lat
ter assumption makes the radius of the edge drop out of the final results.
The full equation for i// given in [8] reduces to a simpler expression if the
normal pressure on the rear slider flank can be neglected. It becomes
^ — [/iexp \i (ir - a - 6.)]/[sino + p cos 8, exp JU(TT - a - 8.)] (6.5)
where a is the complement of the angle of attack (see figure 6.8 (d)); δe is
defined by the equation
cos a exp [-ju(w - a) - /x(l - sin 8.) exp (-/i8.) (6.6)
Equation (6.5) is valid only if S. is greater than the angle δ in figure 6.8
(d). As table 6.2 shows, friction ratios calculated for the sliders with μ =
1.0 and 1.5 have the right ranking and agree qualitatively with experiment
for fi = 1.0.
The forces around a moving slider edge create a normal pressure pn in
addition to the static pressure; the theoretical equation for /?, is (10)
p, = constant (1 - p cot y) exp /i (y - a) (6.7)
372 MECHANICS OF PNEUMATIC TIRES

Block sliding direction


FIGURE 6.6. Contact area of rubber blocks of different heights sliding on poly-methyl-
methacrylate (PMMA). (a) Height 2.0 mm, p - 1.3; (b) Height 10.0 mm, /i - 1.0.
From ref. [7] and
TIRE TRACTION AND WEAR 373

sliding
direction

=0,91

FIGURE 6.7. Trisected rubber block sliding (lop) parallel to the cuts, and (bottom)
perpendicular to the cuts. From ref. [7] and 171].

The angle γ is the angular distance to a point on the cylindrical edge


from the horizontal as indicated in Fig. 6.1 1, which shows the distribution
ofpr under the chisel of figure 6.8 (b) with an angle of attack of 90°.
The integral of pr over the projected contact area constitutes an experi
mentally confirmed "frictional lift". Figure 6.12 reproduces time records
of the time dependence of slider heights above a track of unfilled NR. On
releasing the tractive force, the sliders immediately drop by a well defined
amount and then sink slowly because of creep. As the hit originates mostly
in the rear part of the contact, the chisel of figure 6.8 (b) develops a greater
lift with an angle of attack of 25° than with 90°. The initial drop at 90°
374 MECHANICS OF PNEUMATIC TIRES

(a) (b)
V (c) (d)

FIGURE 6.8. Cross section of metal sliders. The shanks are 1.25 cm square (from ref. [8]).

comes from a transient pressure distribution pulling the leading flank into
the rubber before the tensile force has fully developed behind. The fric-
tional lift is difficult to treat theoretically; it will depend on the track stiff
ness, and the lifts in Table 6.3 are therefore given as percentages of the
static indentation. The cylindrical slider is lifted appreciably only on un-

20 30 40 60 80 100

30 40 60 80 100
LOAD, N

FIGURE 6.9. Friction ratios of the sliders infig. 6.8 on two rubber tracks; A: cylinder; B: chisel
with angle of attack of 90° and C: 25°; D: wedge; v - 0.046 cm/sec, (from ref. [8]).
TIRE TRACTION AND WEAR 375

FIGURE 6.10. Distortion of a rubber surface by a hard, spherical slider (from ref. [9]).

filled NR because a sufficient wrapping angle can only develop on this soft
material. The lift is smallest on poly-butadiene rubber which has also the
lowest friction coefficient.

6.1.4 The Temperature and Speed Dependence of Rubber Friction


In experiments on the temperature dependence of friction, speed must
be kept low so that the frictional temperature rise can be ignored. Figure
6.13 shows the friction coefficient of a gum nitrile rubber on glass as func
tion of sliding speed at various temperatures. The experimental data can
be transformed into a single "master curve" as in figure 6.14 by multi
plying the sliding speeds at different temperatures by a suitable factor [11].
The shift factor is a function of the temperature only and becomes a uni
versal function aT for all polymers if a particular reference temperature Ts
is chosen for each polymer which is related to its glass transition temper
ature Tg by
T, = T, + 50

TABLE 6.2 Effective angles 8, andfriction ratios ^ at various angles ofattack (from [8])

Angle of attack (°) - i.o i- 1.0


90 61.0 73.5 2.06 3.52
70 57.5 70.1 1.68 2.76
25 57.6 66.9 1.15 1.77
376 MECHANICS OF PNEUMATIC TIRES

SLIDING DIRECTION

-10

Prel.

Rcosy

FIGURE 6.1 1. Distribution of the additional pressure pr on the edge ofthe moving chisel slider
(from ref. [10]).

as shown in figure 6.15 for four gum polymers of different glass transition
temperature. The solid curve in figure 6. IS shows the temperature depen
dence of aT as given by the WLF equation (6.12)
logar = - 8.86 (T- r,)/(101.6 +T-T,)
which describes the rate-temperature equivalence of typical visco-elastic
phenoma such as the dynamic modulus [13].
TABLE 6.3 Frictional lift, in percent of the static identation (from [8])

Load NR NR SBR BR
k*f unfilled tread tread tread
cylinder 4.95 10.8 — 2.1 1.3
10.55 14.7 2.9 1.9 —
wedge 4.95 17.2 12.1 15.2 6.9
10.55 16.3 15.5 17.5 9.0
chisel, 25° 4.65 14.7 11.0 10.6 7.6
flank leading 9.91 15.2 15.3 10.9 6.9
chisel, 90° 5.29 42 _ 9.0 O2
flank leading 9.30 4.5 — 13 —
11.28 — 9.0 — 1.6
NR = natural rubber, SBR - styrene butadiene rubber, BR = poly-butadiene rubber
TIRE TRACTION AND WEAR 377

FIGURE 6.12. Fractional lift of the various sliders on a track of unfilled natural rubber (from
ref.fSJ).

The existence of this transform is the strongest possible evidence that


rubber friction is a visco-elastic process. It has also been successfully em
ployed to describe the temperature and frequency dependence of such di
verse properties as the rate and temperature dependence of the tensile
strength [14], the rate and temperature dependence of the critical tearing
energy [15] and the speed and temperature dependence of sliding abrasion
[16] (also section 6.2.4).
On glass, the master curve displays a single maximum rising from fric
tion coefficient values of about 0.2 to about 2.5 on the wavy glass used in
the above experiment or even to over 4 on absolutely smooth surfaces like

-2 0 -4-2 0
LOG SLIDING VELOCITY

FIGURE 6.13. The friction coefficient of unfilled NBR rubber on wavy glass as function of
sliding speed at different temperatures (from ref. [11]).
378 MECHANICS OF PNEUMATIC TIRES

-6 -4 -2024
LOG oTV (V in cm /sec)

FIGURE 6.14. The data offigure 6.13 transformed into a master curve of the friction
coefficient as function of log a,v by using the WLF equation (from ref. [11]).

glass, stainless steel [17] or ice [18] and then falls again to very low values
at very high transformed speeds (aTν values). The rubber acts under these
conditions like a more rigid solid and has a friction coefficient of 0.2 to 0.3.
The master curves can either be referred to a single reference temperature,
say 20°C, in which case the speed at which the maximum occurs is the
higher, the lower the glass transition temperature of the polymer (see fig
ure 6.16 a); or they are referred to their standard reference temperature Ts,
in which case their maxima almost superimpose (figure 6.16 b) and small

o A Styrene Butadiene Rubber


a B Acrylonitrile Butadiene Rubber
- C Butyl Rubber
D Isomensed Natural Rubber

-60 -40 -20 20 40 60 140


T-T..X

FIGURE 6.15. Shift factors log arfor four gum rubbers of different glass transition temper
atures as function of the temperature difference between the experimental temperature T and
the standard reference temperature Tr The solid line is the WLF equation (6.12)
(from ref. [17]).
TIRE TRACTION AND WEAR 379

Referred to
their standard
Reference temp,Ts

-4 2 4-4 -2 0
LOG oTV (V in cm/sec)

FIGURE 6.16. Master curves of the friction coefficient of unfilled NR (solid line) and SBR
(dashed line) compounds on a wavy glass track (a) when referred to 20°C; (b) referred to
their standard reference temperature T,— T,+ 50, ie. - 22''C for NR and +4°for SBR
(from ref. [42]).

differences which do occur can be attributed to differences in the fre


quency at which the loss modulus E" has its maximum. If Vm is the veloc
ity of maximum friction and /„ the frequency of maximum loss modulus,
then
VJ1m - A - 6 X 10-' cm (6.8)
where Λ, a constant of molecular dimensions, is the same for all the poly
mers tested. This close relationship to the loss modulus of the rubber sug
gests that on smooth surfaces an adhesional friction process is operating
which is lin ked to the relaxation spectrum of the rubber. Possible mecha
nisms are discussed in section 6.1.5.
On a rough carborundum paper surface, the master curve again dis
plays a maximum, figure 6.17, but its shape is highly asymmetrical, with
the maximum occuring at a much higher sliding speed. At the speed of
maximum adhesional friction, a hump appears on the friction master
curve on clean carborundum which vanishes if the track is dusted with
magnesia powder. In this case, adhesional contact between rubber and
track is largely replaced by the adhesion between magnesia and rubber,
and sliding takes place between magnesia and track. Incidentally, if fric
tion experiments with magnesia powdered rubber test pieces are carried
out on glass, all speed and temperature dependence of the friction coeffi
cient vanishes and a constant friction coefficient of about 0.2 is obtained
over the whole experimental speed and temperature range.
On the dusted carborundum track, however, neither the absolute value
nor the position of maximum friction on the speed axis are affected by the
presence of the powder, and it has been shown for several rubbers of
widely different glass transition temperature that a constant relation exists
between the speed of maximum friction and the frequency of maximum
loss factor, the proportionality constant being of the same order of magni
tude as the spacing of the track asperities. A second friction mechanism is
present in which energy is lost by gross deformation around the asperities
which has been illustrated in figure 6.10 for a spherical slider. Greater
strains are produced on sharp-pointed tracks where adhesional friction is
380 MECHANICS OF PNEUMATIC TIRES

o
*M
c

1
o
o

-4 -2 6 8 10
log Tv (V in cm/sec)

FIGURE 6.17. Master curve of the friction coefficient of unfilled NBR on a clean 180 carbo
rundum track (fiill line with experimental points)—and when dusted with magnesium powder
(full line). The dotted line shows the master curve of the same compound on wavy glass
(fromref. [17J).

amplified, even when dusted, by the inclined flanks of the asperities, as


discussed in section 6.1.3. In addition to these tangential strains, there are
normal strains whose contribution to dry friction can be neglected. They
are, however, the only source of friction on perfectly lubricated tracks, and
of rolling friction, in Greenwood and Tabor's theory [19]. In either case,
the energy losses are determined by the dynamic properties of the rubber,
the surface of which undergoes periodic strains during the passage of suc
cessive track asperities over it. Mechanical losses can, however, be ex
pected even on an ideally elastic material. As shown later in figure 6.47,
rubber is pulled out of the surface by sharp points and eventually snaps
back, its elastically stored energy being irrevocably lost. Higher friction on
rough than on smooth tracks at low transformed sliding velocities, figure
6.17, is attributed to this affect.

6.1.5 The Sliding Mechanism


The empirical relation between speed for maximum friction on glass
and the loss modulus of rubbers, eq. (6.8), has been interpreted as
indication of a stick-slip process on a molecular scale. Small rubber com
plexes adhere to the track by bonds which are periodically broken by ther
mal agitation and by their share of the friction force; they are later re
formed in an advanced position. The loss modulus E" has its maximum at
a circular frequency ω equal to the reciprocal of the mean relaxation time
of the rubber T
« - I/T - (1/T0) exp - (E/kT) (6.9)
TIRE TRACTION AND WEAR 381

where To is a constant and E is the activation energy. If the making-and-


breaking of bonds is approximated by a cyclic process, the length 6 X 10" '
cm in eq. (6.8) is the forward jump during one cycle. The model has been
detailed by several authors reviewed in references [20] and [21]. We shall
briefly describe here one theory in which the making and breaking of
bonds are considered as differently activated processes [22]. A bond has an
average life /and is re-formed after a time lag equal to the relaxation time
r. Assuming a Hookean force displacement law for the stretched bond, the
friction force F becomes
F = NoM vr(i/r)2/(\ + i/r) (6. 10)
where A',, is the number of potential bonding sites and M, the force con
stant of the bond. If W is the adhesional energy, a newly formed bond re
quires the energy fluctuation (E+ W) to be thermally broken. As the bond
is strained by the force f during relative motion between rubber and track,
the energy barrier is assumed to be lowered by λf where A is a length mea
suring the reach of the adhesion force. The ratio (t/τ) in eq. (6.10) is then
given by

i/r = (kT/\M vr) exp TJ / (e~*ly)dy (6. 1 1)


*i
with TJ = (Jtr/λM ντ) exp - (W/kT).
The integral in eq. (6.1 1) is tabulated as —Ei(rj). The rate dependence of
M has been allowed for by [22]
M - AU 1 - exp - (i/r)}/(i/r) (6. 1 2)
where M , is the limiting modulus at high strain rates. The adhesion en
ergy W has been assumed to be due to van der Waals forces and to be
comparable with the thermal energy kT. The dimensionless graphs in fig
ure 6.18 show the theoretical speed dependence of the friction force for
various adhesional energies. The curves resemble experimental results in
exhibiting maxima. The maxima originate from two opposing effects: the
force at which a bond breaks increases, but the number of bonds decreases
with increasing speed. The theoretical curves are, however, much nar
rower than the curves in figures 6.14 and 6.16. This is in all probability
due to the operation of a relaxation spectrum instead of a single relaxation
time. For quantitative comparison between theory and experiment, refer
to [20] and [22].
The theory fails by predicting zero friction at very low and very high
speeds. Failure at high speeds is understandable because the rubber be
comes glassy, and the friction mechanism changes to that of ordinary sol
ids. The theoretical lack of static friction could conceivably be remedied.
If each rubber complex adhered to the track by several bonds (instead of
only one), it could possibly hold indefinitely a small force by an equilib
rium between bonds broken and made, as long as λf < W. Recent work
with transparent sliders or tracks has disclosed the possibility of a very dif
ferent sliding mechanism which, however, appears to have been estab
lished so far only for relatively soft rubbers [9]. Figure 6.19 shows the con
tact area of a natural rubber hemisphere sliding to the left on a Perspex
track. Stereoscopic observation reveals that the black lines at right angles
to the sliding direction are folds in the rubber surfaces where contact with
382 MECHANICS OF PNEUMATIC TIRES

FIGURE 6.18. Theoretical dependence of adhesion friction on sliding speedfor different


adhesion energies (from ref. [22]).

the track has been lost. These folds move rapidly along the contact from
right to left, ie. from the rear to the front of the contact, while complete
adhesion is maintained between the folds which have been called "waves
of detachment". Similar effects occur in the contact between a hard slider
and a rubber track but the waves move here from the front to the rear of
the contact.

FIGURE 6.19. Waves of detachment in the contact area of a spherical NR slider on


poly-methylmelhacrylale at 0.43 mm/sec, (from ref. [7]/).
TIRE TRACTION AND WEAR 383

The phenomenon has been attributed to an elastic instability in the rub


ber adjacent to the other factional member, its origins being tangential
compressive stresses leading to buckling. These stresses have already been
illustrated in figure 6.10 for the case of a hard slider on a rubber track.
Figure 6.20 reproduces photographs of the contact area of a spherical rub
ber slider with latitudinal markings on Perspex. The distortion of the
markings indicates that the sample has rolled over to a certain extent; this
effect is similar to the consequence of the forces /•", and /•'.. on a flat sample
in figure 6.5 (b). It is also seen that the rubber is extended in the front part
of the contact, and compressed at the rear. As the elastic instability giving
rise to waves of detachment comes from a compressive stress, the waves al
ways run from the compressed to the extended part of the contact under
the influence of the stress gradient.
The deformation of a rubber track by a hard slider was calculated for a
simple model of the rubber; the calculated deflection of the rubber surface
was symmetrical with respect to the centre of the contact, with compres
sion in the front part and extension at the rear [9]. An experimental deter
mination of the deflection by Barquins et al. [23, 24] showed, however,
that it was asymmetrical with a maximum right at the front edge of the
contact, as seen in figure 6.21. The points A and B on the abscissa give the
static length of the contact, and A′ and B′ the length when sliding to the
right. The lower graph in figure 6.21 indicates the compressive strain at
the front and the strain gradient.
With the wave mechanism of sliding, the friction force could arise from
the energy losses due to strain propagation around the folds, and from dif
ferences between the work of unpeeling the rubber on one side of the fold
and the surface energy recovered on making contact on the other side.
Roberts and Thomas [25] have shown that adhesion of rubber to glass is
strongly hysteretic and that this hysteresis appears to be connected with
the conventionally defined hysteresis of the rubber. Applying these find
ings to the case of rubber sliding on glass by means of waves of detach
ment, and assuming no other dissipative mechanism to operate, a simple
equation was derived for the friction force F of a rubber slider on glass. If
γ is the loss in surface energy, w the wave velocity, λ the spacing of the
waves and v the imposed velocity
(6.13)

FIGURE 6.20. Contact area of a spherical rubber slider on Perspex when a) resting b) moving
to the left (from ref. (9J).
384 MECHANICS OF PNEUMATIC TIRES

(b)

Q4

02

-0.2

-0.4

-06
-400 -300 -200 -100 0 100 200 300 400
DISTANCE,*, /im

FIGURE 6.21. Deflection (lop) and surface strain (bottom) produced by a spherical glass
slider (4 mm diam.) on NR. A-B, stationary contact; A '-B' sliding contact (from ref. [24]).

Briggs and Briscoe [26] arrived at the same relation between friction and
loss in surface energy, though expressed in terms of different parameters.
Roberts and Jackson [27], having measured F, w and λ, and having de
termined γ independently, compared the experimental friction force of 7
rubbers with their theoretical values according to eq. (6.13). The differ
ences did not exceed 25%, and were as close as 10% in some cases. Waves
of detachment have also been observed on roughened tracks (sand-blasted
Perspex) though with presumably quite soft silicone rubbers [28]. Surface
energy losses calculated from the results were in reasonable agreement
with values deduced from rolling friction. There is, therefore, strong ex
perimental evidence that the friction force of rubber on hard tracks when
"sliding" with the wave mechanism is accounted for by loss in adhesional
energy. It is not known if this finding also holds for hard sliders on rubber
tracks. The relation between sliding velocity, wave velocity and frequency
of wave formation is, however, known to be different in the two cases. [9]
The wave mechanism shows how rubber can slide on smooth tracks
without abrasion. It can also explain static friction. Static friction is deter
mined by the limiting value of the tangential compressive stress below
TIRE TRACTION AND WEAR 385

which the elastic equilibrium in the rubber surface remains stable and no
buckling occurs. Quite recently, Roberts and Thomas have made experi
ments in which the tangential stress in the contact of a spherical rubber sli
der was slowly increased until the first wave began to form, defining this
force as static friction [29]. Its value increased with increasing strain rate,
this effect becoming more pronounced with increasing hysteresis of the
rubber. At zero strain rate, all curves appear to converge to the same level
of frictional stress, the value of which is about 9% of the shear modulus of
the rubber. This proportionality between buckling stress and stiffness is
what would be expected from the observed instability leading to waves of
detachment.
Whether rubber moves by molecular steps of the kind described earlier
in this section or by waves will depend on which process is energetically
more advantageous. The photograph in figure 6.22 shows the contact area
of a butyl rubber sphere on Perspex. The contact is reduced to parallel
ridges between which the rubber has buckled inward but the contact area
did not change during sliding. This is an example of the first step to wave
formation coming into operation but wave formation itself being energeti
cally impossible.
6.1.6 The Surface Condition of Sliding Rubber
When rubber is set sliding on smooth tracks, the friction force increases
at first and only levels out after a travel of a few centimetres [5]. A con
stant value is reached sooner in repeated runs. The effect is quite pro
nounced with NR and could be ascribed to surface crystallization or, at
least, to an ordering of molecules at the surface. Equilibrium friction is
reached quickly on rough tracks because abrasion counteracts any condi-

5mm

FIGURE 6.22. Contact of a spherical butyl rubber slider on PMMA at 0.43 mm/sec, (from
ref. [91).
386 MECHANICS OF PNEUMATIC TIRES

tioning of the surface. It has been found, however, that sliding on abra
sives produces a directional anisotropy of friction when the samples are
later slid on smooth tracks [30]. An example is shown in figure 6.23 for an
abraded NR tread compound which had already been slid once on glass,
curve 1, gives the friction when sliding in the abrasion direction; curve 2,
after turning the sample through 90°; curve 3, when turned back again.
Unfilled NR and SBR tread showed a much smaller anisotropy.
A surface layer on sliding rubber which differs from the bulk in its
physical properties has been revealed by other means. The modulus E in
eqs. (6.2) to (6.4) for the friction coefficient refers to the rubber asperities
rather than to the bulk rubber. There is a difference between the two be
cause the asperities lose part of their initial stiffness through stress soft
ening during sliding [31]. Black-filled compounds also lose much of their
original electrical conductivity because of breakdown of an internal black-
structure. This phenomenon has served to demonstrate the existence of a
distinct surface layer [32].
Samples mounted on metal holders were pulled over a smooth metal
track whilst the electrical impedance between holder and track was mea
sured with alternating current of different frequencies. The impedance be
came frequency-dependent after previous abrasion, or after passage of a
direct current through a new sample during sliding. The results suggested
the presence of a surface layer of high resistance in which local deforma
tion had practically destroyed the black structure. If a direct current po
tential is applied, the voltage drop is concentrated in this layer and gives
rise to a high field strength, resulting in electrostatic attraction between
sample and track which operates like an additional normal load, with a
consequent increase in frictional force. Figure 6.24 is the result of such an
experiment with an abraded tire tread compound of natural rubber; a
moderate voltage (240 V) increases the friction force by about 50%. We re
fer to the original publication [32] for a more detailed analysis of the re
sults. The thickness of the softened layer could be estimated from the volt
age dependence of the friction force as a few thousandths of a centimeter.

90°

10 20 30 40
POSITION ON TRACK . cm

FIGURE 6.23. Effect ofprevious abrasion direction on the frictional force of unfilled natural
rubber (from ref. [30]).
TIRE TRACTION AND WEAR 387


^^cro'
10.0 240\ 240V
on on
7.3 1 ' f°°^
J L*.
240 of 240 Voff
S 5.0 1
> 10 15 2O 25 30 25 40 4 5 55 60 65
uj 120V
|»5 on
ra '
7.5 ~0^—
12 3V off
•s 0 75 80 85 90 95 100 105 110 IIS 120 I2S 130
TIME.min

FIGURE 6.24. Effect of a d.c. potential between sample and track on the friction of a natural
rubber tire tread sample (from ref. (32J).

F rictional anisotropy and electrical effects appear to be connected in the


case of blackfilled BR. As seen in figure 6.25, application of a d.c. poten
tial increases the friction force to a greater measure than for NR, and fric
tion stays high even after disconnecting the voltage; the frictional ani
sotropy after this treatment is strong.
SBR compounds reinforced with HAF black had too high an initial re
sistance to show electrical effects in our experiments but Savkoor and
Ruyter [33] observed an increase in friction of filled SBR on applying di
rect or alternating current potentials. The authors do not, however, state
the grade of black used, nor the relative magnitude of the effect with a d.c.
and a.c. voltage.

20 30 40
POSITION ON TRACK, cm

FIGURE 6.25. Effect of a d.c. potential between sample and metal track on tire friction of BR
tread rubber, and the following directional anisotropy without a potential (from ref. [32]).
388 MECHANICS OF PNEUMATIC TIRES

6.1.7 Rubber Friction on Lubricated Surfaces


If the sliding speed is kept very low, the friction coefficient of rubber on
lubricated carborundum stone at different speeds and temperatures can be
transformed into a master curve as demonstrated in figure 6.26 for the
gum NBR used in the experiments of section 6. 1 .4 on carborundum lubri
cated with distilled water (34). Comparison with figure 6.17 shows that it
displays all the features of the master curve obtained on dry clean carbo
rundum paper. In particular, the adhesion hump is clearly recognisable. If
a detergent is added, this hump disappears. A similar effect was demon
strated in figure 6.17 on dry tracks by dusting them with magnesia.
Roberts [35] has recently shown that polar substances like soaps can
very effectively prevent direct contact between track and rubber (figure
6.27) thus reducing the adhesional friction component. When a rubber
sphere approaches a glass plate lubricated with distilled water, the water
film tends to conglomerate into globules of water, leaving virtually dry re
gions. A strong polar liquid on the other hand maintains a continuous film
to much higher pressures. It appears, therefore, that friction on wet tracks
is determined by adhesion in the dry regions rather than by normal defor
mation losses; this explains the high friction which can be achieved on wet
surfaces. If dry adhesion is admitted, tangential stresses can also enhance
deformation losses, increasing the total friction.
The total friction will be lower than on dry surfaces, because the pres
ence of water reduces the effective contact area. This mixture of dry and
wet contact is often referred to as "boundary" lubrication. At higher
speeds or higher pressures hydrodynamic lubrication plays an increasing
role (see section 6.1.8). The effects of hydrodynamic lubrication on rubber

Wet sihcone
. carbide with
distilled water

FIGURE 6.26. Master curve ofa gum NBR rubber on a wet 180 silicone carbide stone track
with distilled water as lubricant and 5% detergent added to the water respectively
(from ref. f34J).
TIRE TRACTION AND WEAR 389

. 5

. ,1

04 0-2 02 04 04 02 01 0-4
(mm.)

FIGURE 6.27. Newtons fringes showing the topography of a rubber sphere in contact with a
glass plate (a) with distilled water in the contact region and (b) with a soap added to the water
(from ref. 135]).

friction are best studied with sliders of relatively large radius and contact
area in order to reduce the real contact pressure and thus to obtain hydro-
dynamic lubrication at experimentally manageable speeds.
Figure 6.28 shows results obtained by sliding a steel ball over lubricated
thick rubber tracks at various speeds and temperatures. The results are
shown as master curves in order to exhibit the essential features of the
phenomenon more clearly. The master curves obtained show a rise in fric
tion at low aTν values and pass through a maximum to fall again to low
values at high a, v values. On the left branch of the master curve, the trans
formation follows the WLF equation and curve (a) in figure 6.28 shows
that the absolute values of the friction coefficients are very close to these
obtained on dry glass (dotted line). The maximum occurs at much lower
aTν values than on the dry track with the same rubber. Considerable
spread of the data can occur and it is not clear at present whether this is
due to experimental scatter or whether the maximum friction coefficient
obtainable depends on other parameters not considered as yet. Beyond the
maximum, the transformation shift factors are much smaller than would
have been expected from the WLF equation and are probably mainly due
390 MECHANICS OF PNEUMATIC TIRES

to changes in the viscosity of the liquid with temperature. The temper


ature dependence of the viscosity was allowed for (figure 6.28c) by trans
forming all data to the viscosity rj,, at the reference temperature, ie. instead
of a, v as in dry friction, the shift factor is now

The position of the maximum is further a Heeled by the viscosity of the


lubricating liquid. The lower the viscosity, the higher the transformed
speed of maximum friction (figure 6.28 c). The various polymers have
their maxima in the same order on the aTν axis as on dry smooth tracks
but are shifted to the left, and it appears that the difference in position be
tween them is larger. Indeed, the maximum for the SBR gum compound
was outside the experimental range at lower speeds and higher temper
atures than experimentally available (figure 6.28 b). When increasing the
load on the slider, the maximum is shifted toward higher transformed slid
ing speed (figure 6.28 d). When changing the diameter of the steel slider,
the maximum is shifted toward higher speeds with decreasing diameter
(figure 6.28 e). Finally—figure 6.28 f—the position of the maximum de
pends on the hardness of the compound. The harder NR compound, filled
with 50 pphr of HAF black had its maximum at a higher speed than the
softer NR gum compound. These effects can be explained qualitatively if
it is assumed that hydrodynamic lubrication becomes more effective with
increasing transformed speeds.

(b)
Wit polymer track
sphiricol slider —
t kp
SBR gum

-4-202 -4-202
LOGa-rV LOGaTV

FIGURE 6.28. Friction ofa spherical steel slider on a lubricated rubber track (from ref. [34]).
TIRE TRACTION AND WEAR 391

6.1.8 Hydrodynamic Lubrication


If a plane rectangular hard body slides with a velocity ν over a station
ary plane hard surface and the block is free to rotate around a horizontal
axis normal to the sliding direction, a wedge-shaped film of liquid builds
up between the plane surfaces which exerts a pressure p normal to the slid
ing direction and tends to separate the surfaces of the solid bodies. This
effect forms the basis of sliding bearing lubrication. We refer to Mitchell
[36] for an introduction to this subject. The minimum film thickness h is
given by

h = const. (6.14)

where a and b are the block dimensions parallel and normal to the sliding
direction respectively, TJ is the liquid viscosity and L the load. Of particu
lar interest here is the case of a soft rubber block sliding on a hard surface.
In this case the wedge-shaped film forms even if the surfaces run nomi
nally parallel to each other because the rubber surface will deform under
the hydrodynamic pressure. Roberts [35] has demonstrated this very con
vincingly for a rubber sphere sliding on a plane glass plate with a silicon
oil as the lubricating liquid, as shown in figure 6.29. For this case the aver
age film thickness is given by Archard and Kirk [37] as

h - 0.9 IT, (6.15)

where R is the radius of the rubber sphere and E is the modulus of the
rubber. For dimensional reasons a similar relation should also hold for a
hard sphere of radius R on rubber. For very hard rigid bodies, the lubrica
tion film thickness would be determined by equation (6.14) and for very
soft rubbers by equation (6.15). In the case of rigid bodies, the load con
trols the film thickness. When one body is completely pliable, the modulus
controls the film thickness.
It will be assumed that the factional force F is proportional to the effec
tive dry contact area A,, so that
F = const. AD (6.16)
and that the roughness—micro or otherwise—can be represented by
spheres of equal radius r. If the thickness of the lubricating film is h at the
sliding velocity ν, and h is smaller than r, then the dry area protruding
through the film, i.e. the dry contact area is

(6.17)

where n are the number of spheres in the contact area. Using for the film
thickness first the equation (6.14), the friction coefficient /t can be written
as
AD const. 1 a2br\v \n
(6.18)
392 MECHANICS OF PNEUMATIC TIRES

FIGURE 6.29. Liquidfilm thickness and deformation ofa rubber sphere sliding on aflat glass
plate lubricated with silicone oil. The graph at left bottom showsfilm thickness (ordinale) as a
function of contact diameter (abscissa) (from ref. [35]).

or with equation (6.15)

= const.. —» • const. i-M (6.19)

The speed dependence of ju follows then as


- Cv"2) (6.20)
where C has a different meaning according to whether equations (6.18) or
(6.19) are considered as valid. Most friction data on wet tracks at different
TIRE TRACTION AND WEAR 393

speeds in the speed range of hydrody namie lubrication exist as tire skid
measurements and will be discussed in detail in section 6.4.
Only one instance will be examined here. Figure 6.30 shows the skid co
efficient of a smooth tire on fine polished concrete plotted as a function of
v'/2. The five experimental points were themselves averages of nine indi
vidual measurements. It is seen that a straight line plot results within the
slight experimental scatter. The constant C is determined from the inter
cept on the v'/2 axis as 10.2, or all contact is lost at a speed of 104 km/h.
Actually, before this occurs, other mechanisms intervene, notably viscous
drag and the inertia of the liquid film so that the coefficient of friction
never becomes really zero. From equation (6. 1 8) the constant C is given as

• sec
With a = 25 cm, b = 15 cm, TJ = 66 × 10~° -^—\— , L = kgf; v0
28.9 (m/sec), and C = 0.186, the roughness r becomes 0.23 mm which is
certainly close to the observed roughness of the polished concrete surface.
Considering the transformed speeds at which the peak friction coeffi
cients occur in figure 6.28, equation (6.19) explains at least qualitatively
the shifts due to viscosity, radius of the slider and modulus. The peak sig
nals the onset of hydrodynamic lubrication; though not part of the argu
ment leading to eq. (6.19), its location is expected to be determined by the
same parameters as the descending branch of the curve on the right hand
side of the maxima. At constant load, the friction coefficient given by
equation (6.19) depends only on the dimensionless quantity
Rην/Er2 = constant (6.21)
from which it may be deduced that the displacement of the peak along the
speed axis will be such that:

p
0
p
\ Smoo th tin9 on 1 in«_
wet concre te

COEF ICIENT
SKID
Np
Oi^—oi n^ \\_

X>
\
o
\
X
\
1
1 23456789 10
v^km/hr)"2

FIGURE 6.30. The sliding skid coefficient of a smooth tire on a wet concrete track as function
ofthe square root ofthe sliding speed, (size 175 R 14, load 350 kp, inflation pressure 1. 9 bar).
394 MECHANICS OF PNEUMATIC TIRES

on changing the slider radius Rν = constant


on changing the modulus ν/E = constant
on changing the viscosity TJV = constant
ie. the observed shifts are all qualitatively in the right direction.
The effect of load on the position of the peaks cannot be gauged from
equation (6.21) but according to equation (6.18)
v
— - const, would be expected.
LI

Actually, more detailed equations derived from elasto-hydrodynamic


lubrication of spheres [38] suggest that the film thickness decreases with
increasing load much less than indicated by equation (6.18).
Geyer [39] finds that the load dependence of wet friction is similar to
that for dry friction, i.e., the friction coefficient decreases with increasing
pressure. This is not contrary to equation (6. 1 8). The load dependence of
wet friction can only be gauged if at the same time an assumption is made
on the load dependence of the real contact area AD.

6.1.9 The Friction Coefficient on Ice


Figure 6.3 1 shows the friction coefficient of an NR gum rubber on a
smooth ice track as function of the track temperature at different sliding
speeds [18]. At, and slightly below 0°C, the friction coefficient is low be
cause the track is not a stable solid surface. Melting takes place, firstly, be
cause of the pressure between rubber and track and, secondly, because the
cohesive forces between adjacent ice crystals are lower than the adhesion
between rubber and track. As the temperature of the track is decreased,
the friction rises and friction values obtained at different sliding speeds
and temperatures below - 10°C are transformable according to the WLF
transform as shown in figure 6.32. Because of the limited range of data
available, data at different speeds were transformed to an equivalent track
temperature. The possibility of transforming data to master curves in the
usual way has been demonstrated by Southern [40]. If the track is free
from contamination and polished, very high values of the friction coeffi
cient can be obtained. The ice track appears to behave very much like any
other smooth surface once it has attained sufficient rigidity. Influences of
different molecular structures of the track surface, be it glass, ice or metal,
on rubber friction appear to be of secondary importance.
Figure 6.33 shows the coefficient of friction of four gum rubber com
pounds as function of the ice track temperature at a constant sliding
speed. NR displays a broad maximum with a very high friction coeffi
cient. The friction coefficient for SBR rises only over a very small temper
ature range and falls then. The curve for BR rises over the whole experi
mental range of ice temperatures, and NBR never attains high friction
values. The reason becomes immediately apparent if the curves are plotted
as function of T - Tg as shown in figure 6.34, where Tg is the glass transi
tion temperature. In order to bring out the similarity with the log aTν mas
ter curve, the temperature axis has been plotted in the opposite direction
to normal convention, because high T - Tg values correspond to low aT
values. Since the shape of the master curve is similar for different poly
mers and they differ primarily in their position on the log aTν axis, the fric-
TIRE TRACTION AND WEAR 395

-5 -10 -15 -20 -25 -30


ICE TRACK TEMPERATURE ,"C

FIGURE 6.31. Friction coefficient of an NR gum compound on smooth ice as function of the
ice track temperature at three different sliding speeds. 3 5 cm/sec • S x 1(T' cm/sec O 5 X
lOT2 cm/sec (from ref. [18]).

tion coefficients of various polymers when plotted as a function of the dif


ference between ice-track and glass transition temperatures, become part
of a single curve which is a distorted version of the master curve on
smooth surfaces. BR has its maximum value at lower temperatures than
the experimental range. NR is at its maximum value, SBR and NBR have
their maximum friction at higher temperatures.

s^ O" i smoothice
FCOREIF,/i
OF CITCIEONT
a
A n $0.
a-S
-^" ^-i A ^°
b
b
'oo a in
O

a
^ a
,--
X\ wovy
glass
,^"
—t01o
• .''''
.''
f \
t
\ O
I
1
Data for different si ding < peed . war* 1
transformed occor< ling t 3 the
WLF equation
1
-9 -10 -15 -20 -25 -30 -35 -40 -45 -50 -55
r
ICE TEMPERATURE ,°C

FIGURE 6.32. The friction data in figure 31 transformed to a constant sliding speed of 5 x
Iff1 cm/sec, using the WLF equation and data on wavy glass (from ref. [18]).
396 MECHANICS OF PNEUMATIC TIRES
4.0

-5 -10 -15 -20 -25 -30


ICE TEMPERATURE ,"0

FIGURE 6.33. The friction coefficient on ice offour gum polymers of different glass transition
temperatures asfunction ofthe ice track temperature at a constant sliding speed (from ref. [18]).

6.1.10. Filler Effects and Oil Extension


Gum rubbers, used above to demonstrate simply the elastomer effects
on different surfaces, play no part in tires because they are too soft. Tire
compounds contain invariably a substantial amount of filler, pre
dominantly carbon black, but in some cases also silica or blends of carbon
black and silica. The inclusion of a filler does not affect the WLF transfor
mation of the friction curve; it changes, however the shape of the master
curve considerably, as shown in figure 6.35. On glass, the maximum is pro
gressively reduced as the carbon black is increased, presumably because
the increasing stiffness reduces the real contact area. On carborundum pa
per, the deformation peak is also rapidly reduced in height as the black
filler level is increased; with 50 parts of filler, adhesion peak and deforma

100 90 80 TO 60 50 40 30 20 10 0 -10
T-Tg, «C

FIGURE 6.34. The data offigure 33 replotted asfunction ofthe difference between ice track-
and glass transition temperature. The temperature scale has been reversed in direction to
bring out the similarity with the log aTv — master curve (from ref. [42]).
TIRE TRACTION AND WEAR 397

-a -4-2024
LOG oTV (V in cm/sec)

FIGURE 6.35. The friction master curve on glass of (a) NBR as gum rubber and (b) when
filled with 50 HAF (from ref. [42]).

tion peak have about the same value so that on clean abrasive, the master
curve shows a broad plateau reaching from the speed of maximum adhe
sion friction to that for maximum deformation friction [11]. These findings
have more recently been confirmed by Rieger [41] whose results are repl-
otted in figure 6.36. All curves have extensive plateau regions, except for
BR which does not appear to have reached its adhesion maximum within
the experimental speed and temperature range because of its very low
glass transition temperature.
Synthetic rubbers for tread compounds are now always oil extended
during polymerisation. This enables processing at much higher molecular
weight and improves such physical properties as wear and fatigue resis-

10s 108
10" 10" to" 1.0
OTV (v in cm/sec)

FIGURE 6.36. Master curves for the coefficient offriction of different rubbers filled with
carbon black sliding on emery paper at 20°C (from ref. [41]).
398 MECHANICS OF PNEUMATIC TIRES

tance. Friction, too, is influenced by both filler and oil content. Natural
rubber can be oil-extended by adding the oil with the filler in the mixer. It
saves processing time and confers similarly improved properties on it as
for synthetic rubbers. When compounding to equal hardness or modulus,
it appears that the oil acts as a replacement for rubber, ie. the filler ratio is
virtually unchanged if it is referred to rubber plus oil. Its effects on other
properties, however, such as hysteresis and notably friction are entirely
different from simple replacement.
Figure 6.37 compares the friction master curves of a gum NR with an
oil extended black filled NR compound on a glass track [42]. It is seen by
comparison with figure 6.36 that when oil is added with the black, the fric
tion is much less reduced than by the filler alone. The rubber behaves
more like the gum rubber, although additional black has been incorpo
rated to achieve the same hardness as that of an unextended tread com
pound. Comparing oil extended tread compounds based on NR and SBR
respectively with unextended ones on wet carborundum it is seen from fig
ure 6.38 that, over a considerable range of a,v values, oil extension raises
the friction coefficient for both polymers. It will be shown in section 6.1.12
that this is the range important in tire traction. Since the range of a,v val
ues is limited, cross-overs in tire traction coefficients with temperature and
sliding speed can be expected and indeed have been observed in practice,
(see section 6.4.8). A new filler, silicon-oxide, has recently been added to
some tread compounds to increase friction, notably on icy surfaces. The
effect of silica on the master curve on glass is shown in figure 6.39. Al
though the compounds were a little softer than the corresponding black
filled ones, because the stiffening effect of silica is less pronounced than
that of carbon black, there is no indication from these data that silica does
improve the friction coefficient.

6.1.11 The Effect of Polymer Blending


Blending of elastomers differing in glass transition temperatures and
other physical properties to obtain synergistic effects is common practice
in tire technology. In the present context, its effect on the friction coeffi
cient will be considered but other effects, notably on wear (see section
6.5.2) and fatigue resistance also play an important part. Figures 6.33 and
6.34 demonstrated that on ice at moderate temperatures, only natural rub
ber was capable of a high friction coefficient; for SBR, the temperatures

-7 -6 -5 -4 -3 -2-10 1 3 4
LOG 0TV,(V in cm/sec)
reference temperature Ts« -H1°C

FIGURE 6.37. Thefriction master curve on a glass track ofan NR (a) gum compound and (b)
oil extended tread compound 100 NR/50 oil/70 HAF.
TIRE TRACTION AND WEAR 399

-3 -2 -I 04 3 2-1 0
LOG oTV, (V in cm/sec)

FIGURE 6.38. Comparison of the friction master curves on wet sillcone carbide stone for
unextended NR and SBR tread compounds with the corresponding oil extended mixes.

are already too low and for BR, too high. However, blending the two poly
mers in equal ratio produces the friction coefficient shown in figure 6.40.
Also shown again are the two curves for the pure polymers. It is apparent
that the range of high friction coefficients has been extended considerably.
When attempting to place the curve into the friction coefficient vs. (T -
Tg) plot of figure 6.34 it appears that an effective glass transition temper
ature of about — 80°C would have to be assigned to the blend; this is about
half-way between the glass transition temperature of BR (-108°C) and
SBR (-46°C). This extension of the range of high friction is maintained in
the presence of black filler or a blend of black and silica and with oil ex
tension (18); an example is given in figure 6.41. It is seen that these find
ings not only hold for SBR-BR blends but also for NR/BR blends.

65432101234
LOG QTV (V in cm/sec) referred to their Ts temperatures

FIGURE 6.39. Thefriction coefficient master curve on a glass trackfor oil extended SBR and
NR tread compounds filled with HAF black in comparison with similar compounds filled with
SiO2 (100 polymer/50 oil/70 filler].
400 MECHANICS OF PNEUMATIC TIRES

4.0

-5 -10 -15 -20 -25 -30


ICE TEMPERATURE ,»C

FIGURE 6.40 Thefriction coefficient of a SBR/NR gum rubber blend and of unblended SBR
and NR on smooth ice as function of the track temperature; (from ref. [18]).

6.1.12 The Range of the aTν Values in Tire Skids


Temperature and sliding speed are only independent variables as long
as temperature rise in the contact area is small. At sliding speeds as they
occur during tire skids or even during partial braking, the temperature rise
in the contact patch is considerable. Figure 6.42 shows the temperature he-

>c r _ 60 SBR/40BR-OE

-5 -10 -15 -20 -25 -30


ICE TRACK TEMPERATURE,*C

FIGURE 6.41 Comparison of the friction coefficient of an oil extended NR/BR tread
compound with an oil extended SBR/NR blend as function of the ice track temperature.
Also shown are the curves for OENR and OESBR.
TIRE TRACTION AND WEAR 401

(ween a rubber track and a conical slider the tip of which was formed by a
thermocouple. When the track is dry, very high temperatures may be
achieved. Even on wet tracks a rise in temperature is observed at low
speeds, indicating that dry contact exists between rubber and track. At a
certain speed, the temperature reaches a maximum, falling again at higher
speeds. It is suggested that hydrodynamic lubrication (see section 6.1.8)
becomes effective as the sliding speed is increased and track and slider are
separated by a thin continuous water film. The friction coefficient is then
low and the temperature cannot rise. The speed at which the maximum
temperature rise occurs depends on the load on the slider, its radius of cur
vature and the hardness (modulus) of the track. Knowing the temperature
in the contact area and the sliding speed, the a,v values for different poly
mers can be calculated as function of the speed, as shown in figure 6.43.
At very low sliding speeds, at which the frictional temperature rise is
small, a, v increases because of the increasing speed. As the temperature
rise becomes noticeable, aT decreases progressively as ν increases, even
tually outweighing the speed effect so that aTν passes through a maximum.
It is more quantitatively deduced from the total friction-speed coeffi
cient

(6.22)

where (d/i/dv)r is the slope of the master curve at the working point (ν,T)
and dT/dν is the frictional temperature rise. As d(log aT) dT is always neg
ative, dfi/dν is always smaller than (dfi/dv)T and can easily have the oppo
site sign.
The consequence is that in practical tire skids on wet and dry surfaces

Conical slider DRY TRACK/-


I mm tip radius
2.5 kp load 1 /
300
/ Spherical
¥. II mm $
2.5 kp load
I
§ ZOO

Conicol slider
too- I mm tip radius
\ 2.5 kp load

WET TRACK

10 20 30 40 50 60 70 80
SUDING SPEED, km/hr

FIGURE 6.42 The temperature in the contact area between a rubber track and a conical slider
as function of sliding speed (from ref. [42]).
402 MECHANICS OF PNEUMATIC TIRES

0.01 1.0 10
SLIDING SPEED, km/hr

FIGURE 6.43 log a-fV as function of sliding speedfor NR and SBR, estimatedfrom the
temperature rises offigure 6.42.

the range of a,v values is limited and hence only a small section of the
master curve is realisable. Its value depends further on the difference be
tween the contact and the reference temperature, ie. on the glass transition
temperature of the elastomer in question: The lower the glass transition
temperature, the smaller the aT value; for any given sliding speed and con
tact temperature, the operating point on the master curve moves further to
the left with decreasing glass transition temperature, as is apparent from
figure 6.43 which shows that aTν is lower for NR than for SBR at the same
sliding speed.
When tires slide on icy surfaces, the temperature cannot exceed 0°C so
that the aT ν values are higher by several decades than they are on wet or
dry surfaces and increase with the sliding speed. Whilst on dry or wet sur
faces the operating point on the master curve is on the left of the maxi
mum, on ice it is most likely on the right.
Figure 6.44 shows the master curves of gum SBR and NR on glass, re
ferred to their standard reference temperatures, with the average aTν val
ues operative on wet and icy surfaces marked. It shows that a reversal in
ranking is to be expected when passing from the temperature and speed
conditions (aTν) prevailing during a wet skid to those on ice, and this re
versal is indeed observed in practice. The case is somewhat hypothetical
because the master curves are those for gum rubbers on glass. However, it
has already been demonstrated that a polished ice surface of sufficient ri
gidity behaves very much like a glass surface. Contamination which dra
matically lowers the friction in practice could at worst obliterate com
pound effects if no rubber contact were made at all between track and tire
but it could not reverse the ranking. It is now also clear why the blending
of SBR with BR in tire tread compounds reduces the friction on wet tracks
but increases it on ice.

6.2. Abrasion
6.2.1 Introduction
As has been pointed out in section 6.1.1, sliding on smooth tracks does
not necessarily produce abrasion; generally, abrasion is initiated by the lo-
TIRE TRACTION AND WEAR 403

0-6-4-2024
LOG oTV (Vin cm/tec)

FIGURE 6.44 The friction coefficient ofNR and SBR gum rubbers on glass; the points mark
the average values operative on dry, wet and icy tracks.

cal stress concentrations at the contact between track asperities and rub
ber. The intensity of abrasion depends on the shape rather than on the size
of the asperities. A five-fold increase in the grain size of garnet paper in
creases the abrasion of tread compounds by only about 50 to 70% [5], but
abrasion on silicon carbide paper, grade 180, is 3 1/2 times greater than on
a much coarser aluminium oxide stone, grade 36 [43]. It is also found that
abrasion on different public road surfaces can differ by a factor 4 or 5. We
shall later distinguish between blunt and sharp tracks; this division,
though useful in the discussion of abrasion mechanisms, is not clear-cut in
practice. The ultimate criterion is the effect of antioxidants and the sur
rounding atmosphere on the rate of abrasion which shows that different
failure mechanisms operate on tracks with different surface topology.

6.2.2 The Abrasion Pattern


When a new rubber sample is continuously abraded in the same direc
tion, the rate of abrasion increases at first before reaching a constant
value. At the same time, the rubber surface develops an array of nearly
parallel ridges at right angles to the abrasion direction [44] examples of
which are shown in figure 6.45. A cross-section through these ridge forma
tions, which have been called abrasion patterns, reveals their profile to be
saw-tooth shaped, with the teeth pointing against the abrasion direction
(figure 6.46). During sliding, the teeth are bent backward and expose their
under-cut side to abrasion while the rear side is tucked in and protected.
The teeth wear progressively thinner until their tips are torn off, leaving
only vestiges some of which are seen in figure 6.46. This almost macro
scopic volume loss constitutes a significant addition to the ordinary abra
sion which proceeds on a much smaller scale. When pattern formation is
hindered by periodically changing the abrasion direction, the stationary
level of abrasion is considerably lower than when maintaining the same
direction. A corollary of these findings is that laboratory abrasion ratings
404 MECHANICS OF PNEUMATIC TIRES

FIGURE 6.45 Abrasion patterns produced on NR and SBR tread compounds on different road
surfaces. The numbers give the percentage abrasion of SBR relative to NR (from ref. [5]).

ABRASION

FIGURE 6.46 The profile of abrasion patterns on gum NR abraded on silicone carbide cloth
(upper) and on a worn tire surface (lower picture) (from ref. (44J).
TIRE TRACTION AND WEAR 405

can depend on the nature of the track, as seen from the data noted in fig
ure 6.45.
The initial stages in the formation of abrasion patterns are illustrated, in
somewhat exaggerated form, by the model experiment of figure 6.47 in
which a conical point was dragged across a soft rubber surface with paral
lel markings to show the ensuing distortion. A tongue of rubber is pulled
out of the rubber and eventually snaps back. On tracks with closely spaced
asperities, the upper side of the tongue is abraded before it can retract.
"Figure 6.47 also shows a tear at the root of the tongue which advances into
the rubber. The abrasion pattern thus digs, as it were, its own root and is
self-perpetuating but takes some time to develop. A further observation is
that patterns move bodily along the rubber surface in the abrasion direc
tion; the tears at the root of the teeth must therefore have a forward com
ponent.
y Abrasion patterns appear to be a consequence of the high friction to dy
namic stiffness characteristic of rubber. They are common on laboratory-
abraded samples and are, as expected, more intense on soft than on hard
rubbers. They are also stronger when abraded on blunt than on sharp
tracks because they can then develop before being worn away. Patterns
occur on NR tires but are less conspicuous on synthetic, and practically
absent on oil extended treads. This does not mean, however, that the ini
tial stages of the process shown in figure 6.47 do not operate on synthetic
treads. A recent theory of abrasion described in section 6.2.6 starts from
such considerations.
When patterns are found on tires, their orientation has diagnostic value

FIGURE 6.47 Distortion of rubber surface by a conical needlepoint moving over it (from ref.
406 MECHANICS OF PNEUMATIC TIRES

because it indicates the direction of relative motion between tread and


road.
6.2.3 Load Dependence
The load dependence of abrasion is important in any treatment of tire
wear because the contact pressure, apart from depending on the load, var
ies over the footprint area. Figure 6.48 gives results, expressed as depth of
abrasion per unit sliding distance, for four filled compounds on garnet pa
per, a prototype of a sharp abrasive. The samples had been extracted with
acetone to avoid smearing, and were periodically turned through 90° in
their own plane to prevent abrasion patterns. (Smearing effects are de
scribed in reference [5] and will be discussed here in section 6.2.5). The
proportionality between abrasion and load in figure 6.48 is well within the
experimental error limits. Remembering that friction on sharp tracks is
also nearly proportional to load, figure 6.48 can be taken to mean that
abrasion on these surfaces is proportional to the factional energy dis
sipation. The proportionality factor must involve the relevant strength of
the rubber, and its nature can be surmised as follows: Let the volume of
the rubber particles detacheo during abrasion be <? where a is propor
tional to the contact length between track asperity and rubber; let the dis
tance between consecutive abrasion sites also be proportional to a. Taking
the appearance of needle scratches [45] as a guide, local tensile failure ini
tiates abrasion, and a particle is detached after only a few passages of as
perities over it. As discussed in more detail in section 6.2.6, crack growth is
generally governed by the tearing energy T which, under severe condi
tions, is proportional to the energy density at break of the rubber [46].
Tear propagation is similar in the present case to that of a 'trousers' test
piece for which the tearing energy is proportional to the applied force and
thus is given by the product of normal pressure and coefficient of friction.
Assuming that the effect of other rubber properties on the process can be
neglected, the abrasion depth h becomes, for dimensional reasons:
h = const. (w/U) (6.23)
where U is the energy density at break. Equation (6.23) will be docu
mented in the following section. The abrasion mechanism just outlined is
also thought to work on blunt tracks but the detachment of a particle
needs many more attacks by track asperities than on sharp surfaces. Fail
ure is no longer tensile but becomes fatigue. According to one theoretical
treatment [47] abrasion should then increase more than linearly with in
creasing load. Figure 6.49 gives results obtained with a BR tread com
pound on four different tracks with no obvious sharp points; measure
ments on the carborundum grinding wheel were made on one of its flat
surfaces. Concrete I was a flooring surface, and the other two tracks were
road surfaces. The linearity of the doubly logarithmic plot in figure 6.49
indicate that the pressure dependence of abrasion can in each case be de
scribed by
h = (plptf (6.24)
where n and />.., are empirical constants characterizing rubber and track
[48]. Table 6.4 gives values of n for tracks and compounds which will fig
ure later in the treatment of tire wear. They are all significantly greater
than unity and may thus be taken to confirm that abrasion on such tracks
TIRE TRACTION AND WEAR 407

is, at least in part, due to fatigue failure. This contention will be further
borne out by experimental evidence in section 6.2.5 and 6.2.6.

6.2.4 Temperature and Speed Dependence


Velocity and temperature dependence of the abrasion of unfilled, non-
crystallizing rubber are interconnected in the same way as velocity and
temperature dependence of friction [11]. The family of curves for the ve
locity dependence at various temperatures can be assembled into a master
curve by horizontal shifts, the temperature dependence of which obeys the
WLF equation [16]. As a consequence, the ratio abrasion-to-friction can
be similarly transformed, both friction and abrasion of these compounds
being established as visco-elastic-processes. Figure 6.50 gives abrasion A
(volume loss per unit sliding distance) or A/μ as a function of the trans
formed velocity aTν. The inset shows how the experimental shift factors
follow the theoretical curve. In these experiments the tracks were dusted
with magnesia to prevent smearing by absorbing contaminants. The abra
sion curves are terminated on the right hand side by vertical bars which
mark the onset of stick-slip and which coincide with the occurrence of
maximum friction. The subsidary abrasion maximum in the curve for Bu-

PRESSURE , kg/cm'

FIGURE 6.48 Load dependence of abrasion on game! paper offour different rubbers at 0.66
cm/sec.
SBRwith 50pph HAF black
NR with 50pph HAF black
C NR with TSpph thermal black
D NR with 130 pph Pattinion's activated CaCO,
4M MECHANICS OF PNEUMATIC TIRES

4.0
o Akron Disc
2.0 A Concrete I
A Concrete n
1.0 a Tormac
08
0.6
0.4

i?
0.2
7
0.1
0.08
0.06
0.04

0.02

0.01
0.02 0.04 0.1 0.4 0.6 1.0 2.0 4.0 6.0
PRESSURE, kg/cm2

FIGURE 6.49 The abrasion of a BR tread compound as function of the contact pressure on
different track surfaces (from ref. [48]).

tyl rubber is due to abrasion patterns forming at a particular combination


of friction and dynamic stiffness.
What happens beyond the stick-slip regime is shown by the temperature
dependence of abrasion at a constant velocity, unfilled SBR serving as an
example in figure 6.51. On lowering the temperature, abrasion decreases
until the stick-slip region (and maximum friction) are reached. The tem
perature range of stick-slip is limited; when steady sliding obtains again on
further lowering the temperature, abrasion rises steeply. At the same time,
the abraded surface shows now, in contrast to abrasion at higher temper
atures, pronounced scoring marks like an abraded plastic. The rubber has
become brittle at the rate of deformation imposed on it by the track asper
ities.
The temperature dependence of the abrasion of crystallizing and filled
rubbers follows a similar pattern but the temperature velocity equivalence
no longer holds good. Comparison between the upper and lower graphs in
figure 6.51 shows, however, that a reinforcing black filler reduces the tem
perature coefficient of abrasion at the higher temperature, and also pre
vents stick-slip.
If, as proposed in the preceding section, abrasion on sharp tracks is con
nected with the energy density-at-break, equation (6.23), the temperature
dependence of 1/Ub should be equal to that of A/fi but U must be deter
mined at a strain rate similar to that exercised by the track asperities on
the rubber. This was estimated as about 10000%/sec. for a sliding speed of
1 cm/sec, and an apparatus was built to extend rubber samples up to
break at this speed. The solid curves in figure 6.5 1 are examples of the
temperature dependence of 1/Ub and show that they greatly resemble the
curves. In particular, they have minima at about the same temper-
TIRE TRACTION AND WEAR 409

§s
o
f

*.

w* ^ rN
& — — cs

•5
o

£f*^ Tf
<N —

9
H

Z 00 03
410 MECHANICS OF PNEUMATIC TIRES

-25 0 25 50 75 100 125


2.0

FIGURE 6.50 Master curves for the abrasion loss per unit energy dissipation for three
non-crystallising gum rubbers (from ref. [16]).

-40 0 40 80 -40 0 40 80
TEMPERATURE, "C

FIGURE 6.5 1 Superposition of abrasion loss and the inverse of the energy density at break //
1/4, both as function of temperature (from ref. [16]).
TIRE TRACTION AND WEAR 411

ature below which both curves rise sharply. The minimum of 1/Ub comes
from a maximum in the elongation at break which rapidly decreases at
still lower temperatures. This loss of an essentially rubberlike property
means onset of brittleness, as already inferred from the appearance of rub
ber surfaces abraded in this temperature range.
Figure 6.51 and other results in reference [16] confirm a close correla
tion between abrasion and energy density-at-break. The shift in ordinate
scales to produce the superpositions in figure 6.51 gives the constant in
equation (6.23); we refer to the original paper [49] for numerical values.
No data appear to be available in the literature on the speed and tem
perature dependence of abrasion on blunt tracks.

6.2.5 The Effect of Antioxidants and Surrounding Atmosphere


Brodskii and co-workers [49] carried out experiments on the influence
of an inert atmosphere (N2) on abrasion, mainly on two different tracks,
monocorundum paper and a knurled steel plate. With few exceptions, no
tably Butyl rubber, abrasion was substantially higher in air than in JV2,
and the difference was greater on the blunt steel plate than on the mono-
corundum paper. The same authors also found that certain antioxidants
reduced abrasion although others increased it slightly.
In another investigation on the effect of antioxidants and surrounding
atmosphere on abrasion an Akron abrader with various wheels was used
[50]. Figure 6.52 is the log of the change in abrasion of a protected and an
unprotected NR tread compound on a used standard grinding wheel when
changing the atmosphere from air to N2, and back to air again. (In these
experiments, a 2/ 1 dust mixture of carborundum and Fullers' Earth was
fed into the nip between sample and grinding wheel to counteract smear
ing). Preceding the data in figure 6.52, the samples were run in air until
abrasion had become constant. When continuing in N2, abrasion decreases
gradually before a steady value is reached; the reduction in abrasion of the
unprotected sample is considerably greater than that of the protected

1000 2000 3000 4000


REVOLUTIONS OF GRIDING WHEEL

FIGURE 6.52 Abrasion in nitrogen and air of a NR tread compound with and without
antioxidant on an Akron abrader (from ref. [50]).
412 MECHANICS OF PNEUMATIC TIRES

out*
ft. o, sg
U 01 e

;± K —•

U.
at E
CQ X rl ' c*i O <N ^ o i
" §

>rt p ^; r-; oo ^
wi ^- rn oo oo o

•3-3 z -sz-Sz'sz'Sz

8 S.S S^ E
l
•sf ml
g -c -o a -o 3 -g -o -a E
a!O 5 <
TIRE TRACTION AND WEAR 413

sample. In admitting air again, the first reading is even lower than the fi
nal value in N_. before abrasion rises again to its original load. Final abra
sion of the unprotected rubber is nearly twice that of the protected sample.
Table 6.5 lists equilibrium abrasions for various rubbers and tracks in
air and in A7.,. The metal wheels had knurled rims, and both the dust mix
ture and magnesia were applied to them. For comparison, table 6.5 gives
also results obtained on road surfaces under similar experimental condi
tions. Considering first abrasion of NR compounds in air, all antioxidants
used reduced abrasion, though the effect was small on the road surface
which also produced the highest abrasion rate. On the Akron abrader, the
antioxidants Bayer 4010 and Nonox ZA, which are p-phenylenediamine
derivatives, about halved the abrasion as compared with the non-pro
tected mix. Nonox HFN, which is mostly PEN, gave a lesser improve
ment.
The nitrogen effect is bewildering. In some instances, abrasion is
higher in N2 than in air, against all expectation. The situation can be clari
fied by plotting, in figure 6.53, the difference between abrasion in air and
abrasion in JV., against the abrasion in air. Straight lines are obtained for
any one experimental condition, irrespective of polymer or antioxidant.
This result indicates two essential differences between abrasion in air and
in N2: oxidative degradation leading to an increase in abrasion, and
smearing by degradation products which naturally reduces abrasion. It
had previously been found [16] that magnesia is a better absorber of inter
face contaminants than Fuller's Earth which, however, is a better abrasive
because of larger particle size. Hence, abrasion in N2 (but not in air) is al
ways higher with the dust mixture than with magnesia. The conditions un
der which degradation or smearing occur in air can be deduced from fig
ure 6.53. The effect of smearing outweighs loss in abrasion resistance at
low abrasion levels and loses importance at higher abrasion rates. This

NR , Nonox ZA
o NR,Boyer4010
0 NR, Nonox HFN
NR.no A.O.
SBR, Nonox HFN
BR. Nonox HFN
EPR.noA.O.

-20
10 20 30 40 50 60 70 80
ABRASION IN AIR, mm / 500 revs

FIGURE 6.53 Difference between abrasion in air and nitrogen as function of the abrasion in
air (from ref. [50]).
414 MECHANICS OF PNEUMATIC TIRES

agrees with earlier findings [5]. The straight lines in figure 6.53 nearly pass
through the origin when magnesia is used as dusting agent, indicating that
magnesia virtually prevents smearing.
Figure 6.52 can now be interpreted. During abrasion in air a degraded
surface layer is formed on the sample which, on changing to N2, must be
worn off before equilibrium in N2 is established.
When changing to air again, the abrasion resistance is still high at first
but smearing sets in instantaneously so that the rate of abrasion is quite low
immediately after changing from N2 to air.
The effect of antioxidants and atmosphere on abrasion closely parallel
their effects on fatigue life [51, 52], thus strongly suggesting that this type
of abrasion is due to fatigue failure. Subsidiary experiments have revealed
that humidity, too, affects abrasion, and it appears that part of the N2 ef
fect may come from the gas being dry. This question has not been resolved
yet but could have a bearing on tire wear ratings under different weather
conditions, as may also the nature of the road dust.
Table 6.6 gives road wear ratings of oil extended (75 polymer, 25 oil),
variously protected NR treads filled with 55 pph of HAF black; the mea
surements were made with a two-wheeled trailer [53]. All antioxidants em
ployed are seen to improve wear ratings by up to about 23% as compared
with the non-protected compound. This figure, though useful, falls far be
low corresponding ratings on the Akron abrader but it must be remem
bered that road surfaces vary, and wear will be dominated by the most
abrasive stretches on which tensile failure governs abrasion.

6.2.6 Abrasion by a Blade


Rubber abraded by a blade develops a pronounced abrasion pattern.
Champ, Southern and Thomas [54, 55] could successfully treat this type of
abrasion as a problem in fracture mechanics. As pointed out earlier (figure
6.47), the tongue of rubber pulled out by an asperity or, as here, by a blade
produces a tear at its root. As an established pattern has a semi-permanent
appearance, the vertical component of the tear must equal the depth of
abrasion for one passage of the blade. The movement of the pattern shows
that the tear proceeds in a forward direction into the rubber. If the tear
grows by Δc at the angle /? with the rubber surface, the abrasion depth Δx
per passage is
Ax = Ac sin 0. (6.25)
Crack growth is governed by the tearing energy T, defined as the energy
necessary to increase a tear by unit area [56]; it equals the simultaneously
released energy elastically stored in the sample. The conditions of the ex
periment resemble those of a "trousers" test piece; in the present case, the
tearing energy is then given by
T=(F/b)(\ + COS/6) (6.26)
TABLE 6.6 Road wear ratings on a trailerfor NR treads with different anlioxidant
formulationsfrom ref. [53]

2 pis. 4 pis. 4pts. 4pts.


Anlioxidant None santoflex 13 santoflex 13 Nonox ZA VOP88
rating 91 100 102 112 104
TIRE TRACTION AND WEAR 41S

where F is the force on the blade and b, its width. The crack growth per
strain cycle can be determined independently and ft can be derived from
the motion of the pattern so that Δx can be calculated. The sharpness of
the blade does not enter the result apart from the condition that only one
tooth of the abrasion pattern is deformed at a time.
In the experiments, a razor blade scraped the rim of a rubber disc (Ak
ron sample), tangential force and abrasion being measured under different
loads. The full lines in figure 6.54 give the crack growth as a function of
the tearing energy T, and the points are the abrasion results in the form
A.Y/sin θ for four non-crystallizing, unfilled rubbers. The agreement be
tween the two sets of data is very satisfactory, considering that no arbi
trary factors are used. Difficulties were encountered with normal NR vul-
canizates for which abrasion was much greater than expected from its
crack growth. This divergence has been attributed to strain-induced crys
tallization being more effective in crack growth than in abrasion.

63. Tire Forces on Dry and Wet Road Surfaces


6.3.1 Introduction
** Speed and direction of a self-propelled road vehicle are controlled by
the forces between tires and road. These forces have an upper limit set by

0.1 1.0
TEARING ENERGY, ko/cm

FIGURE 6.54 Crack growth rale and abrasion depth as function of the tearing energy (from
ref. [55]).
416 MECHANICS OF PNEUMATIC TIRES

the available coefficient of friction; once the ratio between horizontal trac
tion and normal pressure exceeds this limit anywhere in the contact area,
local sliding occurs. It is important to remember that sliding friction has
no preferred direction; effective control is therefore lost when sliding ex
tends over the whole contact area. This happens, for example, when the
wheels are locked; then the mass center of the vehicle slides to a stop in
the direction in which it was last travelling when the wheels were locked.
It would appear at first sight that such a contingency arises whenever the
circumferential velocity of the wheels relative to the vehicle differs in di
rection or magnitude from the travelling velocity, because the whole con
tact area should be expected to slide under these circumstances. This rea
soning is, however, valid only for the extreme case of an infinitely stiff
wheel rolling on an infinitely stiff track. Possibly the most valuable prop
erty of an elastic wheel like the pneumatic tire is that it can travel at an
angle to its plane and/or with a velocity differing from its circumferential
velocity without involving the whole contact area in sliding motion.
Strains in the tire surface produced by the ground forces in and around
the contact allow adhesion between tire and road over at least part of the
contact (normally the front part) so that control can be maintained even
when travelling and circumferential velocities differ.

63.2 Simple Slip


The difference between circumferential and travelling velocities is speci
fied by the slip s which is a vectorial kinematic quantity defined as

- (6.27)

where ν is the velocity of the road relative to the wheel axle, and V is the
circumferential velocity of the wheel in the plane of the contact area, also
relative to the axle of the wheel. Δv = v - V is called the slip velocity.
When the wheel rolls in a direction making the angle θ with its plane (side
slip), the slip becomes
5 = sin 8. (6.28)
Pure circumferential slip is given by
s = 1 - V/v = 1 - u/ua (6.29)
where to is the angular velocity of the slipping wheel, and ω0 is the angular
velocity of the wheel when rolling freely at the forward velocity ν.
When braking (ν > V) s is positive, with the maximum value of unity
for a locked wheel. During acceleration (ν < V), s is negative and becomes
negative infinity when the stationary wheel spins. Circumferential slip has
a simple physical meaning. During braking, kinetic energy is consumed in
brakes, tires and road; even when the road is not appreciably deformed, it
takes up a great part of the braking energy as heat. The slip s is the pro
portion of energy lost in tire and road and clearly cannot exceed 100 per
cent. (Positive slips greater than 100% can, in principle, be achieved by
making the wheel rotate against the direction of motion). Similarly, part of
the engine power is lost in the tires during acceleration; s is then the ratio
between losses in the tires and road, and gain in kinetic energy of the ve
hicle. Hence its negative sign and its infinite upper limit.
TIRE TRACTION AND WEAR 417

The strains set up in the contact area of a slipping wheel are illustrated
by the model experiment of figure 6.55 which shows the contact area of a
small, solid wheel on a transparent track [57]. All cases have in common
that a circumferential element of the wheel on entering the contact area
adheres to the track at first. As the element moves further into the contact
area, the imposed slip produces a deflection which increases linearly with
increasing distance from the front edge. This is most clearly seen for side
slip, figure 6.55b, where the deflection is normal to the plane of the wheel
and increases at a rate equal to tan θ. The accompanying surface stress in
creases in the same sense until the local value of limiting factional stress is
reached and the element begins to slide back towards its undeformed posi
tion.
A braking force, figure 6.55c, lengthens an element in the circum
ferential direction before entering the contact area, and the element ad-

(a)

(b)

rtffflflff?^
Id) «4»||||B|^p

FIGURE 6.55 Model experiment to illustrate the conditions in the contact area of slipping
wheels a) free rolling; b) side slip (crab walk); c) braking; d) accelerating. Travelling direction
from right to left (from ref. [57]).
418 MECHANICS OF PNEUMATIC TIRES

heres at first to the track in this state of strain; the deflection of the wheel
increases linearly with increasing distance from the front edge at the rate
of(ν- V)/V= s/(1 - s).
As with side slip, sliding starts (towards the front of the contact) when
the ensuing stresses reach the local frictional stress limit. A driving torque,
figure 6.55d, produces contraction of an element before entering the con
tact region; traction and deflections have the opposite sign to those for a
braked wheel, and the surface elements finally slide out of the contact
area.
The sliding region in the rear part of the contact is, of course, the place
where abrasion occurs. Details of this aspect of slipping wheels wUl be
treated in section 6.5.1.
An exact calculation of the forces on slipping tires presents considerable
difficulties because of their complex structure, but useful expressions can
be derived from a simple model which replaces the tire by a toothed
wheel, it being assumed that the teeth can deform independently of each
other and obey Hooke's law in their stress-strain relationships [58]. The
coefficient of friction must enter the calculations. The discussion in section
1 has emphasized the dependence of rubber friction on load, temperature
and sliding speed, all of which vary in the different parts of the contact. A
constant coefficient of friction can nevertheless be assumed in a first theo
retical attempt without grossly violating physical reality.
The deflection of the model is zero outside the contact area. The tan
gential force ts per unit length of contact developed when the model rolls
with side slip is shown in figure 6.56a. Because of the postulated validity of
Hooke's law, the traction increases initially at a rate proportional to the
stiffness k, of the wheel
t, = Jt,x tan 0 (6.30)
where x is the coordinate of the surface element relative to the undistorted
wheel. The limiting tangential traction ftp is indicated by the semi-oval
curve; p is the assumed normal force distribution per unit length. When ts
= μp at the point x = X, sliding begins, and ts in the rest of the contact is
given by
t, = IV- (6-31)
The shaded area represents the sideforce due to the slip sin θ.
A diagram similar to figure 6.56a would depict the force distribution tc
due to circumferential slip, with eq. (6.30) being replaced by eq. (6.32)
*)|. (6.32)
The stiffness kc is considerably greater than ks. Sideforce S, braking
force B and driving force A are formally given by the same type of ex
pression.
If the normal pressure distribution is taken to be elliptical along the
contact area, and constant across it, the calculation yields eq. (6.33) in
which F stands for S, B or A:
2c 2c
(6.33)
1+c2

/, is the normal load, and c is defined by


TIRE TRACTION AND WEAR 419

-Limiting Traction Shaded Area Gives


Total Side -Force

FIGURE 6.56 a) horizontal traction in the contact area of a wheel with side slip; b) traction
resulting from simultaneous side and circumferential slips /i is the limiting friction;
* the shaded area gives the side force in both cases.

c = T- tan 6 for side slip (6.34)

and

c=— ~ 5) f°r circumferential slip (6.35)

with a = length of the contact.


It is easily seen that eq. (6.33) reduces at low slip to
•'small slip ' (6.36)
small slip "A, ni,. 1 1 slip (6.37)
The coefficient of friction has disappeared from eqs. (6.36) and (6.37)
because the contribution from the sliding region in the contact area has
become negligible, and only the wheel stiffness matters. The linear rela
tion between force and slip reflects Hooke's law. At large slip and/or a low
coefficient of friction F tends asymptotically to the value juJL.
Figure 6.57 shows the theoretical dependence of F/fiL on c. It is in
structive to compare the two components making up the force F, ie. the
contribution from the adhesion and sliding regions which have also been
plotted in figure 6.57. The adhesion component, which dominates F at low
values of c, reaches a maximum at c = 1/V5, and is outweighed by the
sliding component when c exceeds about 0.82. What must be borne in
420 MECHANICS OF PNEUMATIC TIRES

mind is that the great contribution made to F by the sliding region of the
contact at large values of c is directionally controlled only by the remain
ing adhesion in the front part of the contact patch, because of the direc
tion-insensitive nature of sliding friction already mentioned.
Figure 6.56 shows that the resultant of the lateral traction making up
the side force does not pass through the center of the contact but is dis
placed to the rear by a distance called the pneumatic trail, resulting in a
couple, the self-aligning torque T, so-called because it tends to reduce the
slip angle. Under certain conditions, particularly on wet roads, the self-
aligning torque can become negative at large slip angles; it then tends to
increase the slip angle (see section 6.3.4). The self-aligning torque 7 for the
toothed model wheel is given by

T 2 (6.38)
3 c2)2

where c is defined by eq. (6.34).


At small slip, eq. (6.38) reduces to
.., 1
(6.39)

Following Gough's suggestion [59], the side force has been plotted in
figure 6.58 as function of T, both quantities being expressed in dimension-
less form. The graph reproduces the essential features of experimental
curves, in particular the maximum of T.
The driver has four items of information to help him steer a vehicle:
The change in steering wheel angle, the change in direction of the vehicle,
the force on the steering wheel due to the self-aligning torque and the ac-

0.2 04 0.6 0.8 10 12 1.4 16 18 20

FIGURE 6.57 Force F on a slipping wheel as Junction of the variable C according to eqi.
(6.33) (6.34) and Us two components.
TIRE TRACTION AND WEAR 421

0.01 002 003 004 005 006 0.07

FIGURE 6.58 Theoretical relation between side force and self-aligning torque according to
eqs. (6.33) and (6.38).

celeration due to the side force acting on the vehicle. Detailed experiments
have shown that accurate handling in rapidly changing situations is only
possible if proportionality exists between the force on the steering wheel
and the acceleration perceived by the driver, ie. proportionality between
self-aligning torque and side force. A decrease in the rate of self-aligning
torque with steering wheel movement signals the onset of critical condi
tions to the driver [60]. To keep perfect control of the vehicle the side force
should be between 0.65 and 0.75 p.L when the self-aligning torque reaches
its maximum [61].

6.3.3 Load-slip Equivalence [62]


The only variable determining the ratios S/μL and T/aμL in eqs. (6.33)
and (6.38) is the quantity c which is defined in eq. (6.34) and takes in both
slip and load. The theoretical curves for S/fil and T/aμL vs. c should
therefore describe both their slip and load dependence. To test this con
tention, logarithmic plots of (S/L) as a function of tan θ measured at dif
ferent constant loads were superimposed on the theoretical curve. The
necessary vertical displacement to produce a fit gives the friction coeffi
cient fi,, and the horizontal shift gives the quantity

so that the cornering stiffness 1/2 k#- can be determined. A result of this
procedure is shown in figure 6.59, using data for a bias tire measured on a
flat bed machine [63]. An independent check on the validity of this load-
slip equivalence is afforded by the initial slope of the S vs. θ curves which
equals 1/2 ksa\ according to eq. (6.36). The experimental points in figure
6.59 follow theory quite closely, apart from data for the lowest load (600
422 MECHANICS OF PNEUMATIC TIRES

Ibs) for which the experimental curve is somewhat steeper at large slip
than expected. The values 1/2 ksa2 determined by the two different meth
ods differed by no more than 3%.
Similar considerations apply to the self-aligning torque. The data for
this particular tire could also be assembled into a universal curve, al
though less satisfactorily than for the side force, and the horizontal shifts
needed for superposition differed seriously from those for the side force
[62]. It will be shown in section 6.3.8, however, that both side force and
self-aligning torque of radial tires obey theory over a wide range of the
variables.

6.3.4 Composite Slip


Tires are often required to transmit simultaneously lateral and longitu
dinal forces to the ground, a common instance being driving wheels in a
curve. Lateral traction ts and longitudinal traction tc act then jointly in the
contact, giving the total traction /
C,2 + (6.40)
and sliding sets in when t equals μp. As a consequence, either traction dif
fers from what it would be if only its own associated slip operated. Figure
6.56b illustrates the effect of longitudinal slip on the side force. The point
X is shifted forward as compared with the case of only one of these trac
tions acting. Furthermore, the factional force in the sliding region is now
taken up by both the lateral and longitudinal stiffness of the tire so that
only part of the sliding friction is available for the side force. The shaded
area giving the side force is much smaller than the corresponding area in
figure 6.56a for side slip only. These considerations apply to braking and
driving forces which reduce the side force equally on the basis of this
simple reasoning.
The tire forces have been worked out with the assumption that the stiff
nesses ks and kc remain constant, and that the tractions ts add tc have the
same ratio in the sliding as in the adhesion region of the contact.

LOAD (Ibs)
o 1800
a 1400
o 1000
• 600

FIGURE 6.59 The quantity S/pL as function of c for a bias tire, size 7.50-15 (data from ref.
[63]).
TIRE TRACTION AND WEAR 423

Theory gives the following equation for the side force S as function of
simultaneous side slip sin θ and circumferential slip ,v:
2a 2a
( }
where
j = cos 6 - V/v
and
IT
8 cos 0 - s

The ratio between side force and either braking or accelerating force is
S/B - S/A - (k, sin ff)/(kfs). (6.42)
The self-aligning torque becomes

T = y /tLa ^ , *, si

(6.43)

When kjk, > 1 and s > 0, T becomes negative at large slip because the
longitudinal tractions, being laterally displaced, make a negative contribu
tion to the torque. As an example of the theoretical dependence of S and T
on the longitudinal force, figure 6.60 gives the results for a tire with 1/2
kff - = 1.46 kgf, L = 400 kgf, μ = 1 and kjks = 3, rolling at a slip angle of
6°. S decreases with increasing absolute value of the longitudinal force
which progressively monopolizes the available friction. The reason for the
negative torque at large braking forces has just been stated. Figure 6.61
shows experimental results [63]. Figures 6.60 and 6.61 are similar, but the
experimental side force actually increases at small braking forces. This ef
fect has not always been observed and appears to depend on tire structure.
The problem of composite slip has been studied in more detail by Pacejka
[65].

6.3.5 Cornering
The cornering power of a tire is not given by the side force S itself but
by its component in the direction of the centrifugal force C, that is by S
cos θ. The inset in figure 6.62 shows the equilibrium of forces on an other
wise free wheel; A represents the drag force in cornering. Ordinarily, the
difference between S and S cos θ is negligible but can become significant
in tight turns with low road friction. The relation between S, S cos θ and θ
is best visualized by a polar diagram in which 5 (or S/L) makes the angle
θ with the direction of C; the ordinate gives S cos θ directly, as in figure
6.62 which demonstrates the effect of road friction on the cornering power.
All curves coincide at very small slip because S is dominated there by the
424 MECHANICS OF PNEUMATIC TIRES

125

1.00

0.75

0.50

025

FIGURE 6.60 Theoretical dependence of side force and self-aligning torque on longitudinal
forces; 1/2 k^a2 - 11.46 L - 400 kgf,• n = 1; k<k, = 3; slip angle 6".

tire stiffness (eq. 6.36). The curves then diverge, go through maxima at slip
angles which decrease with decreasing friction, and finally become circles
of radius μL centered at the origin. The slip angle for maximum cornering
power is a critical quantity because the radius of the curve travelled by the
wheel increases on further increasing the slip angle. The usual response of
turning radius to steering is reversed. This contingency is likely to occur
on slippery or icy roads and severely impairs handling of the vehicle.

, 1000

-1200 -800 -400 0 400 800 1200


FORCE , Ibs

6.61 Experimental dependence of side force and self-aligning torque on longitudinal


forces (flat belt machine).
broken lines: II •«
solid lines: « - 4° (from ref. [63]).
TIRE TRACTION AND WEAR 425
Figure 6.63 shows the forces on a highly idealized rear-driven vehicle in
a curve. The diagram, though showing a 4-wheeled car, has actually been
worked out for a one-dimensional vehicle for which near- and off-sides
coincide, and the mass center coincides with the geometrical center, self-
aligning torque and rolling resistance have been neglected. The subscripts
/ and r refer to front and rear, and n and p are the directions normal and
parallel to the chassis.
The vehicle has to counter the centrifugal force, marked as the vector C,
with its components Cn and Cp. To balance the centrifugal force, the
wheels have to make the slip angles θf and θr with the instantaneous veloc
ity ν in order to develop the side forces Sr and Sf at right angles to their
planes. The sum of their components in the direction of C cannot by itself
balance C; an additional force has to be supplied in the direction of the
chassis in order to offset the drag forces Df and Dr of the slipping tires.
This force is the source of the energy loss incurred when negotiating the
bend and is a measure of the wear accompanying cornering. It is, of
course, delivered by the engine. The front wheels, being free to rotate,
cannot take up a circumferential force, and can only supply the side force.
The driving force D for both front and rear wheels has therefore to be pro
duced by the rear wheels, bringing about circumferential slip in addition
to side slip. If front and rear wheels carry the same load and are inflated to
the same pressure, the rear wheels need then a larger slip angle than the
front wheels to produce a given side force, for reasons discussed at the end
of the preceding section.
The outcome of a braking torque in a curve is sketched in figure 6.64 for
a free wheel. Figure 6.64a shows the tire on a circular path of radius R
before braking; the immediate effect of a braking force B, figure 6.64b, is
two-fold. As the tire makes the slip angle θ with the instantaneous driving

0.1 0.2 0.3 0.4 0,5 0.6 0.7 0.8 0.9 1.0
(S/U sin 8
FIGURE 6.62 Polar diagram of S/L as Junction of 9 on roads with different friction. The
ordinate gives the theoretical cornering power S cos 9.
426 MECHANICS OF PNEUMATIC TIRES

Centre of turn
6
FIGURE 6.63 Equilibrium offorces on a rear-driven vehicle describing a circular curve in the
stationary state. Rolling resistance, wind resistance, load transfer and self-aligning torque have
been neglected.

direction, B produces a component B sin θ in the direction of the centrifu


gal force so that the existing cornering power S cos θ no longer suffices to
balance C. Secondly, the braking force reduces the side force at the slip
angle θ to, say, S1, exacerbating the situation. The radial component of the
tire forces B + S1 gives the new maximum centrifugal force C1 which can
be balanced and which is smaller than C. At a given velocity the radius of
the turn must therefore increase to
(C/C)R. (6.44)
At the same time, the velocity of the vehicle is reduced by the decelerating
force
-m dv/dt = B cos 0 + S' sin B. (6.45)
A vehicle not only begins to describe a wider curve but it also starts with
an angular momentum around its center of gravity which is determined by
its moment of inertia and its angular velocity around that axis before
braking. Conservation of momentum thus makes the vehicle turn around
TIRE TRACTION AND WEAR 427

its vertical axis with respect to its new travelling direction. After this in
stantaneous effect, the velocity decreases; if the steering is not corrected,
the vehicle begins to describe a curve of smaller radius, and the risk of los
ing control is increased.
The idealised model in figure 6.63 can at best show the nature of the
problem. In reality, the centre of mass is never at the geometrical centre so
that, for this reason alone, side force at front and rear are different during
cornering, and the front and rear wheels have to travel at different slip an
gles.
Front wheel drive vehicles are interesting in this respect. Their mass
centre is generally nearer the front; because of the driving torque the side
force of the front wheel is reduced at a given slip angle. This requires large
slip angles and the vehicle has a tendency to move on a larger radius than
the driver envisages. In a curve, the vehicle is less likely to turn around its
vertical axis because of decreasing angular velocity. Finally, the rear
wheels develop their full side force capability since they are free-rolling.
This corrects any tendency of turning more than intended in a curve, and
the car feels more stable in all critical situations.

6.3.6 The Effect of Load Transfer


The model used in figure 6.63 for the calculation of the cornering forces
excluded load transfer from inner to outer wheels in a curve. This problem
can, however, be treated easily without having recourse to a specific tire
model. If two tires on a given axle are required to produce a total corner
ing power 2 C under the combined load 2 L, then
2 C = C0 + C, = OS0 + S,) cos 0 (6.46)

Centre of turn
before broking
(o)

New centre of turn


(b)

FIGURE 6.64 Effect of braking on a single wheel during cornering.


428 MECHANICS OF PNEUMATIC TIRES

where both wheels are assumed to have the same slip angle; the subscript 0
and i refer to outer and inner wheels respectively. Let the load on the outer
wheel be L + ΔL, and L - ΔL on the inner wheel. By series expansion
C0 = C + (SC/SL)^^, AL + (8C/8ff)M
C, = C + (8C/8L\L^L,,L AL + (8C/8ff)M (6.47)
where Δθ is the slip angle adjustment necessary to keep the vehicle on a
given curve. The subscripts in eq. (6.47) indicate the load range of the dif
ferentials.
Combination of eqs. (6.46) and (6.47) leads to
+ 2(SC/8ff) = 0. (6.48)
Expansion of the expression in square brackets gives as final result
Ad = - ((82C/8L2) (8C/8ff)] (AL)2 (6.49)
which shows that Δθ increases with the square of the load difference.
Eq. (6.49) can be applied to the toothed wheel tire model used before.
With S given by eq. (6.33)
[c/(l +c2)]3sindcosd AL
(6.50)
c2)2] - (775/4/iL) sin2d L

where c is again given by eq. (6.34). Eq. (6.50) is greatly simplified for
small slip, ie. the range in which S can be taken as proportional to the slip
angle. The expression becomes then
Ad = (rf/4) (S/fiL)2 6 (AL/L)2. (6.5 1 )

6.3.7 Side Force and Self-aligning Torque at Varying Slip Angle


It has so far been tacitly assumed that the side force and self-aligning
torque appear instantaneously with the slip angle. In addition to the expe
rience of inertial forces already referred to in sect. 6.3.2, there are time
delays between a varying slip angle, side force and self-aligning torque
which are essential handling characteristics.
The simplest tests are carried out on the test rig by oscillating the tire
around a vertical axis through the center of the contact with a sinusoidally
varying slip angle, and measuring amplitude and phase angle of side force
and torque.
A simplified theoretical approach to the problem assumes small slip an
gles so that the tire can be taken to adhere to the track over the whole con
tact; inertial effects are neglected. Because of the latter assumption, the tire
axle can be considered stationary, and the road to change direction peri
odically. Figure 6.65 gives the lateral deflection y of a tread element dur
ing its passage through the contact. If the element enters the contact at the
time f = t0, its coordinate x at the time i is given by
x + a/2 = V(t - /„) (6.52)
dx= Vdt
where 'a' is again the contact length and V, the circumferential velocity.
TIRE TRACTION AND WEAR 429
During dt, y changes by θ dx, where θ is the instantaneous slip angle, and
by x d8, where dli is the change in slip angle during dt.
Hence
dy - 0 dx + dO. (6.53)
Continuing the calculation for the toothed wheel model of the tire, the de
flection at the front edge of the contact is zero (as sketched in figure 6.65)
and y is obtained as a function of x and / by integrating eq. (6.53) from t,,
to t. With
0 sin ut (6.54)
the result is

y-8m x sin ut + — sin u t- (6.55)

The first term signifies bodily oscillation of the track around the center
of the contact; the second term describes a wave with the velocity V and
the wavelength X
A - 2irV/u. (6.56)
The side force S is derived from eq. (6.55) as
/+«/2
y dx m sin (ut + <p,) (6.57)
«/2

where op, is the phase angle of 5.


Writing
au
(6.58)
~V

(6.59)
9- = - p/2. (6.60)
Similarly, the self-aligning torque T becomes

T=kJ yxdx = Tm sin (ut + <pr) (6.61)

Loterol Deflection

FIGURE 6.65 Change in lateral deflection in the contact area at varying slip angle.
430 MECHANICS OF PNEUMATIC TIRES

with
Tm - \/Ukcf -* {72 + f + 6 [p1 + (p1 - I2p) • sinp
+ (5p2- 12) cos p]}1'2 (6.62)
3p (1 + cos/>) — 6 sin/>
tan <f>, (6.63)
p2 — 6 (1 — cos/?) + 3psmp
Velocity and frequency dependence of amplitudes and phase angles are
shown in dimensionless form in figures 6.66, 6.67, 6.68. Figure 6.66 reveals
one drawback of the tire model used here. When the velocity V is low
enough for the contact length to be a multiple of the wave length λ, S van
ishes. Thus, there is an infinite number of zeros between V = ωa/2ir and V
= 0. The reason is that the tread of the model has no bending stiffness and
can be distorted into a close sinusoidal shape which a real tire cannot sus
tain. Similarly, the self-aligning torque and its phase angle begin to oscil
late at these low speeds, Tm/(1/12 ka3) converging to unity, and <p7 to zero.
The phase angles become zero and the amplitudes constant at high speeds.
Experimental side force results in figure 6.69 agree qualitatively with
figure 6.66, apart from the difficulties at low speeds. Data for the self-
aligning torque in figures 6.70 and 6.7 1 allow a more detailed comparison
with theory because of the intricate value of the curves. Theory predicts
the sharp maximum of T at a speed at which the phase angle changes from
positive to negative values, and also the minimum of <i , at about 30%
higher velocity. In contrast to theory, <p, changes sign again at a higher
speed. This divergence could be ascribed to dynamic tire losses which are
not considered in the calculations.
Quantitative comparisons are necessarily limited in view of the primi
tive tire model but it is of interest to estimate the contact length 'a' from
the experimental data. Choosing for this purpose, as a well documented
quantity, the velocity at which the phase angle <pr first becomes zero, a
contact length of 87 cm is obtained, which is about 5 times its actual value.
It is known, however, that it takes about one revolution of a tire to reach

-10O«

FIGURE 6.66 Theoretical side force, and phase angle <t>, between side force and slip angle, as
a function of reduced frequency v/2a w.
TIRE TRACTION AND WEAR 431

FIGURE 6.67 Theoretical self-aligning torque as a function of reduced frequency v/2a a.

equilibrium conditions after changing the slip angle whereas a travelling


distance of only one contact length suffices for the model used here.
A more detailed treatment of the response of tire forces to variation in
slip angle is given later in this book.

6,3.8 The Influence of Road Conditions on Side Force and Self-Aligning


Torque
Figure 6.72 shows the response of side force and self-aligning torque to
a sinusoidally varying slip angle. Since the equilibrium side force-slip

FIGURE 6.68 Theoretical phase angle tfr between self-aligning torque and the slip angle as
a function of reduced frequency vf2d ox
432 MECHANICS OF PNEUMATIC TIRES

150 I I
f« 0.85 Hz
a 125

olOO
cc
euj 75
o
<0
90

25

20 40 60 80 100 120 140


SPEED, km /hr

FIGURE 6.69 (a) Experimental speed dependence of the sideforce at a slip anglefrequency of
0.85 Hz.

angle characteristic is virtually linear over the slip angle range covered,
the response is also sinusoidal although with a phase delay. Its magnitude
depends on frequency and speed of the tire as outlined in the previous sec
tion; it will also depend on constructional details of the tire. The self-align
ing torque response is far from sinusoidal because the static characteristic
passes through a maximum within the large experimental slip angle range
used here. This is demonstrated in figure 6.73 which shows the non-linear
static self aligning torque-slip angle characteristic as constructed from the
response to a sinusoidally varying slip angle. When calculating phase
angle delays from a harmonic analysis of input and response function, this
non-linearity must be borne in mind, otherwise very misleading results
may be obtained.
Side force or self-aligning torque plotted as a function of a varying slip
angle give hysteresis-like loops which degenerate into ellipses if either side
force or self-aligning torque depend linearly on the slip angle, as recently
shown by Schilling [66]. It is also possible to plot the side force directly
against the self-aligning torque. Such diagrams are shown for four slip
angle amplitudes and three frequencies in figure 6.74. Again, if both side

i i i I l I i I 1 1 i i
Phose Difference between Slip Angle and Sideforce

10 20 30 40 50 60 70 80 90 100 110 120 130 MO 150


SPEED, km/hr

FIGURE 6.69 (b) Experimental speed dependence of the phase angle between side force and
slip angle at a slip angle frequency of 0.85 Hz.
TIRE TRACTION AND WEAR 433

40 60 80 100 120 140


SPEED, km/hr

FIGURE 6.70 Experimental speed dependence of the self-aligning torque at a slip angle
frequency of 0.85 Hz.

force and self-aligning torque were linear functions of the slip angle, el
lipses would result.
Figures 6.75 and 6.76 show equilibrium side force and aligning torque
amplitudes of a steel-belted radial tire on wet surfaces. Carpet diagrams
are used to combine the load and slip angle dependence. It was shown
above that side force and aligning torque at small slip angles theoretically
do not depend on the friction coefficient but only on tire construction fea
tures, so that vehicle control is not unpaired by wet road conditions at
small slip angles. Inspection of figure 6.75 shows that this is certainly true
for the side force at high loads. The slope of the side force-slip angle
curves are the same on both surfaces. At the low load, however, the side
force on gravel is smaller than on the sharp asphalt under similar testing
conditions.
Figure 6.77 shows the transformation of the side force and aligning
torque carpet diagrams of figures 6.75 and 6.76 into the universal curves
discussed in section 6.3.3 for dry surfaces. At low V values, the side forces
for the gravel surface fall below the expected values but otherwise it is
possible to fit all data to the universal function. From the shift factors, the

uj 80
Phase difference between self aligning torque and
i §60
l§40

; 20
I ° 20 40_>r--T 80 100 120 140
SPEED, km/hr
jg20
' £ 40

FIGURE 6.71 Experimental speed dependence of the phase angle between the self-aligning
torque and the slip angle at a slip angle frequency of 0.85 Hz.
434 MECHANICS OF PNEUMATIC TIRES

FIGURE 6.72 Time dependence ofsideforce and self-aligning torque at a sinusoidally varying
slip angle.

effective friction coefficient and the tire stiffness, 1/2AX can be calculated
from eqs. (6.33, 6.34). 1/2/fX was also obtained from the slope of the side
force-slip angle curve at the origin. Finally, the contact length a can be ob
tained from the ratio of the slope of the aligning torque-slip angle curve to
the side force-slip angle curve at their origins. It can also be calculated
from the transformation and can be compared with the directly measured
values. All are shown in table 6.6.
The most relevant quantity in this context is the friction coefficient: on
asphalt it exceeded unity, ie., it was much larger than the values associated
with sliding friction coefficients on wet road surfaces. On gravel it was
much smaller than on asphalt but again larger than the sliding values nor
mally obtained on this type of track. They correspond more closely to the
peak values obtained in braking experiments (see section 6.4). If the peak
occurs at 15% slip, then the sliding speed at 30 km/h is, on the average,
0.15 × 30 = 4.50 km/h. In the trailer experiment it was theoretically even
smaller 30 X tan 40 = 2.0 km/h. Under these conditions, hydrodynarnic
lubrication effects are small in both cases.
Directional effects, which might be expected because the tread pattern
reacts differently to hydrodynamic effects in the lateral and longitudinal
directions, are negligible because the sliding speed is low in both cases. A
correlation between the peak sliding value in braking experiments and the
TIRE TRACTION AND WEAR 435

— Side force slip angle


(characteristic
Side force

FIGURE 6.73 Construction of the "static" side force ami self-aligning torque—slip angle
characteristic from data obtained at a sinusoidally varying slip angle.

maximum obtainable side force has, indeed, been reported in the litera
ture [67].
Gengenbach [68] has determined maximum side forces and aligning
torques under diverse conditions of water level, tread patterns, speed and
road. These data reflect essentially what has been said above on the hy-
drodynamic effects. When the sliding speeds become sufficiently large, hy-
drodynamic lubrication takes over, and the side force obtainable is re
duced in the same way as the peak coefficient in braking experiments.
Figure 6.78 shows the phase angles between side force and slip angle,
self-aligning torque and slip angle, and side force and self-aligning torque
obtained on two surfaces for different slip angles and loads. They were
calculated from the first harmonic of the experimental curves. They are
mostly negative ie., force or moment are trailing the slip angle and the side
force is trailing the self-aligning torque.
The phase angle between side force and slip angle becomes less negative
with increasing slip angle and more negative with increasing load. The
phase angle between self-aligning torque and slip angle shows a similar
trend. At low loads and large slip angle the sign reverses, the self-aligning
torque appears to lead the slip angle. This effect is partly due to the highly
non-linear response of the self-aligning torque to slip angle variations, but
436 MECHANICS OF PNEUMATIC TIRES

s .IF A NGLE .« M ^L TL DE
• 0
tV €0
u
5 400
-^

a 200 \ ^
E /I I //
y // /
0.3 0 f
if
pnn / // /
(ikk.
AAf>

400
?r\r\
. -.
N

I1
' //
t.O

200
0
ffj 1
/1
/ '/
/
//
I k.

400

400
•V,
^

200 ' ~~\ ) '


n / / ) /
3.0 0
L 1
/ / j/
S / / /
200 f
J ( •*-*
^
400
-10-50 5 10 -10-50 510 -10-50510 -10-50510
SAT.mkgf

FIGURE 6.74 Side force vs. self-aligning torque at a sinusoidally varying slip angle for four
different slip angle amplitudes and three slip angle frequencies.

Surface
400 400 .Wet
Surface: Wet Gravel Asphalt

4'. .700
.300 300

: 200 200

100

700500300 700 500 300


LOAD.kp LOAD.kp

FIGURE 6.75 Side forces of a radial tire on wet road surfaces as function of load and slip
angle. Results were obtained with the ETDC Uniroyal tireforce measuring trailer; tire size 1 75
SR 14.
TIRE TRACTION AND WEAR 437

20 Surfoce: Wet Gravel 20 Surfoce: — 1 700 •


Wet
Asphalt

15 15

i 10

300 300
700500300 700500300
LOAD.kp LOAD.kp

FIGURE 6.76 Self-aligning torque accompanying suie forces in figure 6.75.

it can also he real as in figure 6.7 1 where at high speeds the SAT begins to
lead the slip angle. The most important phase angle is that between side
force and self-aligning torque because this is what the driver senses when
steering. Normally it is negative, ie., the side force follows the self-aligning

-0.5

-1.0

-1.0

-2.0

-1.0 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6


LOGC

FIGURE 6.77 Transformation of the data from figures 6.75 and 6.76 according to eqs. (6.33)
and (6.38).
438 MECHANICS OF PNEUMATIC TIRES
WET GRAVEL WET ASPHALT
-2
-4
-6
•6 .300
-10 TOO
-12
-14
4T,3OO~
10

900
TOO
n
300-
900
,700
3OO MO 700 I • 0.
-to - 900 4*
TOO'
TOO 0
TOO,

-zrf 300

900

FIGURE 6.78 Phase angle differences between (a) side force and slip angle, (h) self-aligning
torque and slip angle and (c) side force and self-aligning torque on two different wet roads
and with different slip angle amplitudes and tire loads [Kg]. Tire 1 75 SR 14, inflation pressure
1.9 bar, frequency: 1 c/s.

torque. It becomes more negative with increasing slip angle (time delay
becomes larger) and less negative with increasing normal load. At 1 ° slip
angle there is little difference in phase angle between the two surfaces at
all loads. At the larger slip angles, however, the phase angles are much
larger in absolute values on the wet gravel than on wet asphalt. Although
these phase differences constitute only very short time intervals (10° corre
sponds to 30 m sees.) they contribute to the "handling characteristics" of a

Travelling Distance
STEERING
LATERAL
AC ELERATION ANGLE(°)
WHEEL FORCE.kgf, S
(%g)_ in
OOOOl
O _T -
/
-
-
/
/ \
\\ // \\ -
-
-/ / V —
*
X „/ 1
V S

-
f N / \
-/,
= /
v_ ^ \ S
-
X \ y/ 5
_/
_ . \^~ N^s— -

TIME, sec —-

FIGURE 6.79 Steering angle movement, side ways acceleration andforce on steering wheel of
a car driven through a slalom course at a nominal speed of 50 km/h with a pylon separation of
20m.
TIRE TRACTION AND WEAR 439

tire which may loosely be defined as the safe, or otherwise, feeling which a
skilled driver has during extreme steering manoeuvres.
Figure 6.79 shows part of a record of steering wheel movement, steering
wheel force and sideways acceleration in a car driven through a prescribed
slalom course. It will be noticed that both the force on the steering wheel
and the lateral acceleration are delayed relative to the steering wheel
movement. A time delay exists also between sideways acceleration and the
force on the steering wheel. These time delays depend not only on the ex
perimental conditions such as speed, slalom amplitude and surface condi
tion but also on the tire construction and inflation pressure and correlate
with subjective tire handling ratings by drivers. Together with maximum
attainable sideways acceleration, they are used to assess the merit of par
ticular constructional features of tires.
63.9 The Influence of Tire Construction on Side Force and Self-algning
Torque
The most important difference felt by the driver between bias, bias-bel
ted and radial ply tires with either textile or steel belt, is due to their differ
ent side force and self-aligning torque-slip angle characteristics. Figure
6.80 compares a bias tire and a steel belted radial ply tire of similar size.
The slope of the side force-slip angle curve is smaller for the bias tire than
for the radial ply one; under given cornering conditions (speed and curve
radius), the radial ply tire will negotiate the curve with a smaller slip
angle.
The self-aligning torque for both tires passes through a maximum at
about the same slip angle; maximum value and the change with slip angle,
however, is larger for the radial ply tire. Although the non-linear region
between side force and self-aligning torque sets in more abruptly (see

)lo jor
ft Belted —
7200
$/ Steel
Radial Ply
Ti e
too ,/
/

2468
/
02468
SLIP ANGLE C) SELF ALIGNING TORQUE , mkp

FIGURE 6.80 Side force and self-aligning torque as function of slip angle of a bias tire and a
radial ply lire of similar size under the same load and inflation pressure.
440 MECHANICS OF PNEUMATIC TIRES

Gough diagram in fig. 6.80) for the radial ply tire, the fact that it occurs at
a higher side force, gives the driver added safety in cornering.
The phase difference between side force and self-aligning torque is more
negative (side force trails the torque) for the radial ply tire than for the di
agonal one. This does not necessarily mean that the driver feels this as un
desirable. A steep change in side force with slip angle and a very short
delay time leads to over-reaction and makes the vehicle react "nervously"
to small steering corrections.
Of course, the tire force characteristics of radial ply tires, ie. side force
amplitude, self-aligning torque amplitude and time delay between them
are further influenced by constructional details, such as stiffness of the car
cass and sidewall, and belt construction (number of belt plies, cord angle,
cord density, width, coating compound of the belt and many more). Fi
nally, the tread pattern design and the tread compound influence the tire
force characteristics through modulus and friction coefficient, as demon
strated in figure 6.81 which shows a Gough diagram for two radial ply
tires of the same construction under different loads. One tire had the full
tread depth, the other had its tread depth ground to half the original
height. The values plotted are the amplitudes in a dynamic experiment
with sinusoidally varying slip angle. Table 6.7 shows the phase angles be
tween the side force and self-aligning torque for the two tires. At small slip
angles the tire with the lower tread height has a smaller time delay be
tween side force and self-aligning torque whilst at large slip angles this dif
ference disappears. There is even a tendency for the tire with the lower
tread profile to show a slightly larger phase delay.
The differences in the handling characteristics which result from the re
duction of the tread height can be read from figure 6.8 1 and table 6.7. For
the same side force, the self-aligning torque is larger for the tire with the
lower tread height. At low loads the maximum is reached at somewhat
higher side forces for the tire with the lower tread height whilst at high slip
angles a rather sharp maximum occurs in the self-aligning torque at lower
side forces. In extreme steering situations, the tire with the lower tread
height would handle less well than the one with the full tread height.
When effecting small corrections for wind disturbance or during over
taking manoeuvres the tire with the lower tread height reacts more directly
at slightly increased steering wheel forces, which the driver may well judge
to be an asset.

TABLE 6.7 Phase angles between sideforce and self-aligning torque ofa tire with 9 mm tread
height and the same tire with 4.5 mm tread height

1° slip angle amplitude 6° slip angle amplitude


load tread height tread height
(kp) 9mm 4.5 mm 9mm 4.5mm
300 12.2 6.1 31.2 33.9
400 g.l 3.5 19.2 19.8
600 5.7 1.8 10.2 12.6
frequency 1.0 c/s
speed 100 km/h
tire size 155 SR 13 steel belted radial
inflation pressure: 2.1 bar
TIRE TRACTION AND WEAR 441

400

Treod Height
of 4. 5 mm
I I
2 4 6 8 10 12 14 16
SELF ALIGNING TORQUE , mkp

FIGURE 6.81 Gough diagrams for 9 mm and 4.5 mm tread height at different loads and slip
angle amplitudes.

6.4. Braking and Traction of Tires


6.4.1 Experimental Methods
Two distinct methods have become widely accepted as standard ways of
measuring the skid resistance of tires: (a) the front wheel braking test and
(b) the locked wheel trailer test. In the first method the rear wheel brakes
of a test car are disconnected, the test car is driven at a prescribed speed
onto the skid track, the car is declutched and the front wheel brakes are
applied with steadily increasing pressure so that the front wheel brakes
lock after an interval of about 1/2 sec. They are allowed to remain locked
for one further second. During this time the deceleration and the speed of
the vehicle are recorded [69]. Because of the relatively small decrease in
speed during the braking experiment, the dependence of the skid coeffi
cient on speed can be determined readily by changing the approach speed.
A wide range of speeds can be covered. Since only the front wheels are
.locked, the side force acting on the rear wheels stabilises the direction of
the car up to the highest speeds. The deceleration can be converted into a
skid coefficient, denned as ratio between braking force per tire and normal
reaction, taking account of the mass of the vehicle, the load distribution
and the load transfer during braking. A further correction allowing for the
tilt of the vehicle during braking is readily applied [70]. The skid coeffi
cient n is given by

(6.64)

where F is the frictional force, Rf is the normal reaction on the front


wheels, and c1 and c2 are two vehicle constants which have to be deter
mined for the test vehicle. /* is the measured deceleration in % of the
earth's gravitation g.
When braking the car, its rolling resistance also affects its deceleration.
442 MECHANICS OF PNEUMATIC TIRES

This is easily taken account of if the car is declutched and allowed to free-
roll for a short period (1/2 - 1 second) before the brakes are applied. If
the deflection of the recorder trace is measured from the trace correspond
ing to the free-rolling condition the rolling resistance is eliminated. If this
is done for every measurement, any variation in rolling resistance due to
changes in speed, wind conditions, etc. can be removed.
In the second type of experiment, a trailer is pulled behind a towing ve
hicle and the tow bar pull is measured after the brakes of the trailer wheels
have been applied. Either one or two wheeled trailers are used. Both have
a number of disadvantages over the car skid test. The tow bar pull is the
force between towing the vehicle and the trailer, and any acceleration or
deceleration of the car during the skid experiment enters as an error. This
can also be overcome to some extent by declutching the towing vehicle
and allowing it to free-roll before applying the brakes of the trailer. The
rolling resistance of the towing vehicle will have to be determined in a sep
arate experiment. Also, the range of speeds which can be covered with the
trailer is limited since it tends to oscillate around the towing hook at high
speeds. Finally, more power is required for the trailer-towing car combi
nation than for a test car to achieve the same result, and investments are
correspondingly larger. The skid coefficient is calculated from the tow bar
pull F and the load L as
H-c#'[l + C4Lf] (6.65)
where c3 and c4 are trailer constants determined by its dimensions and u! =
F/L.
6.4.2 The Skid Coefficient Versus Time Curve
After the brakes are applied the braking force rises with time, passes
through a maximum and falls to a lower value which remains almost con
stant, as seen in figure 6.82. Experiments have shown that the slip at the
maximum is about 15%; the wheels lock shortly after the maximum value
is passed. In some cases, notably on very smooth surfaces, at high speeds,
or if the surface is covered by a deep layer of water (more than 1 mm) the
sliding value remains constant over a considerable time interval. In other
cases, however, particularly on rough surfaces with a thin layer of water
and at low speeds, the sliding value of the skid coefficient increases with
time (figure 6.83). This is also observed at low speeds on dry surfaces. As
the sliding speed decreases at the same time, this could be interpreted as
an increase in sliding friction with decreasing sliding speed. This negative
speed dependence of the skid coefficient is discussed in some detail below.
It is by itself not sufficient to explain the time dependence of the sliding
value of the skid coefficient during a single skid experiment. If the experi
ment is repeated with different approach speeds and both skid coefficient
and tire sliding speed (forward speed of the test vehicle) are continuously
recorded it should be possible to assemble data for different approach
speeds into a single curve of the skid coefficient versus instantaneous
speed, independently of any variation of sliding friction with time. This is
often not the case as seen in figure 6.84 which shows the skid coefficient of
a tire on a wet asphalt track as a function of speed. The different symbols
represent results with different approach speeds. Generally, the tire which
has skidded for a long time has a higher skid value at a given speed than a
tire shortly after the skid has started ie. closer to the approach speed: The
TIRE TRACTION AND WEAR 443

TIME, seconds
FIGURE 6.82 Time record ofthe deceleration ofa car in a locked wheel braking experiment.

conditions in the contact area change not only with speed but also with
time. The origin of this is not completely understood but is thought to be
brought about by abrasion. This assumption is supported by the fact that
it occurs mostly on sharp surfaces and if it occurs it does so on dry and wet
surfaces.

6.43 The Slip Dependence of the Skid Coefficient


The skid coefficient versus time curve is a distorted version of the slip
dependence of the skid coefficient. As the brakes are applied the slip be
tween tire and road rises until it reaches its limiting value when the wheels
lock. A longitudinal slip is difficult to control experimentally because of
the existence of a maximum in the friction-slip curve. Less energy is re
quired to maintain a high slip than a lower one so that the tendency is to
wards high slip, ie. locked wheels. As shown in section 6.3, the side force
in lateral slip, and the braking or accelerating force in longitudinal slip,
are independent at low slip of either load or limiting frictional force, and
depend only on the stiffness of the tire. The initial slope of the force-slip
curve is, therefore, the same for all road surfaces and conditions. This is
shown in figure 6.85 for two tires on two wet surfaces [71].

Constant
Speed
Driving

TIME , seconds

FIGURE 6.83 Deceleration vs. time record showing a case in which the sliding friction
increases with time.
MECHANICS OF PNEUMATIC TIRES

10 20 30 40 50 60 70 80 90 100
SPEED, km/hr

FIGURE 6.84 Locked wheel deceleration on wet asphalt as a function of instantaneous speed.
Different symbols correspond to different approach speeds.

The maximum observed on both dry and wet surfaces is not explained
by the theory discussed in section 6.3. Two effects are probably respon
sible. First, the hydrodynamic lift on a wet surface increases with increas
ing slip velocity and reduces the available friction. At the same time, the
temperature in the contact area rises, also reducing the available friction.
The temperature effect is the only one operative on dry surfaces. It is par
ticularly severe at very high slip values; hence the final drastic drop in fric
tion when the wheels lock.
Figure 6.86 shows side force-slip angle curves obtained on dry and wet
surfaces with different friction coefficients between tire and road [72]. For
the bias tire, the initial slopes are the same, independent of the road condi
\ tions except for the curve obtained on "light rain after a dry spell". The
very low friction coefficient is probably due to a thin layer of mud and wa
ter which forms an effective lubrication film. A modern radial ply tire on
two wet surfaces of very different friction coefficients shows, however a
small difference in slope, noticable even at 1° slip angle. On wet gravel
complete sliding sets in around 6° slip angle, with a maximum friction co
efficient of about 0.5 whilst much higher values are obtainable on the wet
sharp asphalt.

6.4.4 The Speed Dependence of the Peak and Sliding Value of the Skid
Resistance
The peak value decreases with speed. The same is true if the sliding
value, measured a constant time after reaching the maximum, is plotted as
a function of the speed at that time. This decrease is due to the hydro-
dynamic pressure which builds up in the contact area between tire and
road. The main influences affecting this have been outlined in section 6.1.8
on lubricated friction. They appear here again. Several equations have
been proposed over the years to describe this speed dependence. Maycock
TIRE TRACTION AND WEAR 445

300
175 SR 14 rodiol ply
1 9 bar/410 kplood

!200

9
in
100

6 8
SUP ANGLE, (°)

FIGURE 6.85 Slip dependence of the brakingforce coefficients B/L of 5.00-16 tires on (a) wet
fine textured asphalt and (b) on wet polished gravel carpet, from ref. [71], and theoretical curves
according to eq. (6.33).

[69] first used the statistical approach and described the speed dependence
simply by a linear regression equation
- ji = b(v - v) (6.66)
where the introduction of the mean value ji allows a comparison with skid
coefficients of different tires directly if the mean speed v is the same in all
experiments; otherwise v has to be adjusted to a common value which

0.8

(b)

10 203O405O60TO8O9OIOO

FIGURE 6.86 Side force as function of slip angle for two different tires on several surfaces
(from ref. 190]).
446 MECHANICS OF PNEUMATIC TIRES

should be well within the experimental range. The general approach


H = fr - bv (6.67)
is not really permissible because to obtain /t,, a considerable extrapolation
of the results is necessary.
If the sliding value of the skid coefficient is plotted as function of speed
on a doubly logarithmic plot, the data can be described by two linear rela
tionships with widely different slopes, as shown in figure 6.87 for several
different tread patterns of radial tires on wet asphalt with a water film
thickness of about 0.3 mm. The whole range can be described by the fol
lowing equation

(6.68)

where /x, and Vc are the skid value and speed at which the slope changes
abruptly. Practically all experimental data known to the authors can be
described by the above relation, including Maycock's original data [69]
[73] which appeared to decrease only linearly with speed. The exponent n
assumes, of course, different values above and below the apparent kink.
Both approaches are, however, without physical foundation. When re-
examining the data on the basis of equations (6.18) and (6.19) it is found
that they describe the data accurately over the whole speed range, as
shown in figure 6.88 for the peak and slide coefficient (from ref. 73). The
constants in the equation are capable of physical interpretation. In its gen
eral form
H - jUoO - cv'/2) (6.69)
iio is the dry friction coefficient which is a function of speed, temperature
and load as explained in section 6. 1 of this chapter. The constant c is con
cerned with the hydrodynamic lubrication parameters and contains the
roughness of the track, the viscosity of the lubricating liquid, the dimen
sions of the slider, modulus and/or the load, depending on the sophisti-

SUDING SKID COEFFICIENT

FIGURE 6.87 Sliding skid coefficient on wet concrete for three commercial radial ply tires.
Each point is the average of nine determinations.
TIRE TRACTION AND WEAR 447

12.5

FIGURE 6.88 Skid coefficient for peak and sliding conditions as Junction of v'".

cation with which the lubrication film thickness is calculated. The model
on which it is based is a simplified view of reality but it is easily seen that
refinements are possible.
6.4.5 Tread Pattern and Surface Rougness Effects
Figure 6.89 shows skid data on a wet concrete track for a smooth, a lon
gitudinally ribbed, and a cross-ribbed tire. Each experimental point is the
average of nine individual determinations. The change in the extrapolated
speed at which p. would become zero according to equation (6.69) when
varying the tread pattern depends on the dimensions of the pattern ribs or
blocks in the contact area parallel and perpendicular to the sliding direc
tion, and on the load on each block.
The difference in this speed between smooth and longitudinally ribbed
tires agrees closely with the changes expected from eq. 6.69. The critical
speed for the cross-ribbed tire, however, is lower than expected from the
theory. The curves for all three tires should converge to the same /to for dry
friction. The calculated values are 1.43 for the smooth tire; 1.23 and 1.21
for the longitudinal and cross-ribbed tires respectively. This is as close as
can be expected in this type of experiment.
Figure 6.90 shows the speed dependence of the braking force coefficient
of a smooth tire on four different surfaces [69], plotted on a linear scale. In
this case the rate of water application was kept constant so that the film
thickness depended on the drainage of the surface. If it is smooth and
close textured, such as fine polished concrete, the water film is thick with
no asperities protruding from it. The braking force coefficient for locked
wheel braking is therefore low for all speeds. Fine cold asphalt is also a
close textured surface but has small sharp asperities; the skid coefficient is
therefore high at low speeds when hydrodynamic effects are small but it
decreases rapidly with speed as the thick water film becomes more and
more effective.
Bridport gravel and quartzite are both well drained surfaces so that the
448 MECHANICS OF PNEUMATIC TIRES

Cross-ribbed tire 60%


net /gross
Longitudinally ribbed tire
8% net/gross
o Smooth tire

67 8 9 10 11 12 13 14
SPEED, v, km/hr

FIGURE 6.89 Skid coefficient under sliding conditions for a smooth, a longitudinally ribbed
and a crown ribbed tire.

water film is thin and the speed dependence of the braking force coeffi
cient is also small. Bridport gravel consists of smooth pebbles with no mi
cro roughness. The thin water film cannot be easily broken even at low
speeds so that the braking coefficient is low. Quartzite has the same micro
roughness but the individual stones have many sharp edges which pene
trate the water film giving rise to high braking force coefficients. Maycock

Polist ed Cc ncrete

Slide
"•V-
»• .
20 30 40 50 60 70 20 30 40 50 60 70
SPEED, miles/hr SPEED, miles/hr

Grovei
Peok
1 —«-»-
-*•*-«
-*-, n
*-*-.
Slide — •

20 30 40 50 60 70 20 30 40 50 60 70
SPEED, miles/hr SPEED (miles/hr)

FIGURE 6.90 Speed dependence of the sliding skid coefficient of a smooth tire on four
different road surfaces (from ref. [69]).
TIRE TRACTION AND WEAR 449

[69] also showed that tread pattern effects are largest on the close textured
surfaces and are smaller on the well drained surfaces. Road surfaces and
tread pattern therefore complement each other in reducing the lubrication
in the contact area with increasing sliding speeds. Good drainage and mi
cro-sharpness of the asperities helping to create high local pressures are
the major contributions of the road surface. Differences between tread
pattern design are most pronounced on smooth close-textured surfaces
which easily form thick water layers during rain, and they disappear
largely on well drained sharp surfaces.

6.4.6 The Effect of Siping


It was shown in section 6.1.2 that slitting a block-shaped test piece in
the direction of sliding creates additional contact area and increases
thereby the coefficient of friction. This requires that the pattern blocks are
rigid in the direction of sliding and pliable normal to it. This is not the
mechanism which is generally suggested for the action of sipes. It is rather
assumed that during partial braking they deform so that their edges wipe
the surface. They are also supposed to transport water from the contact
area during partial sliding. Maycock's data [69] summarized in table 6.8
indicate that their effect on the average friction coefficient is larger for
sliding friction than for the peak value. They also tend to reduce the speed
regression. In this case their effect is larger for the peak than for the sliding
value of the braking coefficient. Kienle [67] suggests that edge sipes in
contact with the grooves can form additional drainage channels. Since
sipes in Maycocks tire, however, did not communicate with the grooves, it
is likely that the sipes interrupt the pressure build up of the lubrication
film in the contact area of this longitudinally ribbed pattern. This may
also create additional local contact area. If employed in block designs, it
would probably have the opposite effect because sipes would weaken the
block and the larger resulting deformation would tend to reduce the total
contact area and lower the friction. Because the sipes combat lubrication,
their effect increases with sliding speed, reducing the speed regression co
efficient of both peak and sliding value on all but the coarse textured, well
drained surfaces on which tread pattern effects are small in any case.

6.4.7 The Effect of Tire Construction


It is often assumed that the better skid resistance of radial ply tires on
wet road surfaces is associated with the more open tread pattern which has
become possible because of the higher wear resistance associated with this
type of tire. Meades [73] has shown that this explanation is insufficient
since radial ply tires with the same tread pattern as diagonal tires have a
higher skid resistance on wet roads for peak as well as sliding friction; this
is the case for both winter tire block patterns and for most cases of longitu
dinal rib design with sipes, as shown in table 6.9. The level of improve
ment depends on the type of road surface. On well drained coarse surfaces
it is small, on close textured surfaces with their thicker water film it is
larger. It is also larger for the peak value than for the sliding value and the
speed dependence, particularly of the sliding value, is less for the radial
ply tires of both pattern than for the corresponding cross ply tires. It is
known that the tread pattern grooves of diagonal tires tend to close more
in the contact area than those of radial ply tires. It is also likely that the
430 MECHANICS OF PNEUMATIC TIRES

1
^** h
O g <N
""
q
m
q
m
^ |a
£ o
4
1 o
y

8 M
1a s -£>
3
1a 8
S!
1
S-o
^
\ * § *O
^
CN OS
* <*
3 •3o -8
1
•Sc
5
.5
h
8 a
eg
•H
~^
00
(S
O
(N

9c o

J S

|
w
c
i I
Q
§ xd ^

ti"55
•«
S
If g
I •i -
3 0
-C

1v

| f 1.0 I
•g- S M.
'"
•s-
1ff 1
•c

1 §«5 1"*
00
5
M
J
frict coef icien speed
of
ratio regres
u (M coefncient
[:
-J O
m
1
TIRE TRACTION AND WEAR 451
tread pattern blocks remain more rigid during sliding if supported by a
rigid belt rather than by a textile carcass, thus maintaining better contact
with the road surface.

6.4.8 The Compound Effect


Several investigations [69] [74] have shown that the ranking of tread
compounds on wet surfaces is largely independent of the test conditions,
as shown in table 6.10. It depends neither on the type of track surface, nor
significantly on the sliding speed in a locked wheel skid, and peak or slid
ing values also rank compounds in the same order. The pendulum skid
tester of the British Road Research Laboratory correlates closely with the
skid ranking of different polymers [75]. Finally, the ranking correlates
with the hysteretic properties of the tread as, for instance, measured with a
pendulum resilience meter [75]. The ranking may reverse only in skid ex
periments on ice. The result of interaction between sliding speed and tem
perature, as discussed in section 6.1.4 and 6.1.12 is that the friction coeffi
cient on wet and dry surfaces, measured under similar testing conditions,
whatever they are, always rank the polymers according to their glass tran
sition temperature. This is the reason why the pendulum skid tester of the
British Road Research Laboratory ranks tread compounds correctly, too.
Since the interval between operating and glass transition temperature de
termines the hysteretic properties, a correlation is generally obtained be
tween the skid resistance rating of compounds and simple hysteretic mea
surements like rebound resilience [75].
In addition to the choice of particular polymers or polymer blends to
achieve a high skid resistance on either wet road surfaces or on ice or to
some extent on both, oil extension is used to raise the skid resistance. In
this case the glass transition temperature is not affected but the coefficient
of friction is raised, as discussed in section 6.1.10, figure 6.38, in the range
of aTν values of the master curve relevant for skidding on wet and dry
roads; thus the skid ranking is improved compared with that of the unex-
tended polymer.
At low log aTν values, the coefficient of friction of oil extended rubber
falls below that of the unextended rubber in a range which could obtain in
tire skids. In this case reversals in ranking would be expected at high con
tact temperatures, ie., at high sliding speeds. Indeed, a higher speed regres
sion coefficient has been reported in the literature [74] for oil extended
rubbers than for unextended ones.
A reversal of compound ranking is observed when braking tires on wet
and icy tracks as shown in figure 6.91 for skid pendulum tests and in table
6.1 1 for braking tests with NR and SBR tread compounds. This also fol
lows from the friction master curves discussed in section 6. 1 .4. As the tem
perature in the contact area cannot be above the freezing point at any
speed rise the operating point moves towards the right into the descending
branch of the master curves. On this part the highest coefficient of friction
is obtained for the rubber with the lowest glass transition temperature.
This is at least true for Nitrile rubber, SBR and NR.
The glass transition temperature of BR is so low that very low temper
atures on ice are necessary for it to reach even its maximum friction value.
Its operating point could well be on the left side of the master curve, even
on ice.
432 MECHANICS OF PNEUMATIC TIRES

si ss 2i
— rO fS ^ oo —
till 1 1 I I I

1 Q O
?323
O O

i i T i T 1 1

I a
M - -3
OO OO

oooooooo
IN OO OO Csl

I
O f>
OO t-;
IN *»;
OO •»
™" IN
r*>. t~;
C4 *O
o o o>
*
do do do do

a 88S88?i8°822?5S!
— vi O ^' Q «rl Q —• O *n O-- Of^ -•
£ ii ii r i T i ii+i+i i
*

5*3 SS 88 «2
OfOfiOfOOOflO— o— —
II II I I + I I I + I I I

.S--*
> -C 02
^;IN *fiN r-_ rn IN— spf*^ ^;IN r-^»r> *r
do do do do do do do d
J

^^,^^,J^SSI
doododddoodo'ddd
•S
.r
1
t-8
"™ V«-S -3-3
V V«-S -S-o
v U-S
u •— -S-S
u —
"M D. un O, "wS O,"!A O. "w Q. 7S C. ~
e
I
g.l
O «- &

« I 1 "i 1 II
1 8 » B I a: =
• i 18 1-i I-a S 8
.a i
TIRE TRACTION AND WEAR 453

ja
1 S S 8
i
ofspe
effect
the
(c) .-Si
over sli
fou
averaged peak
and
+ices

SOmph
s cn
s 1

11 sia*
|
a «n
a £ 00
8

•2 s -0
8
41 1
8
-S

1g 1
"aa T3
> r- 5 8
1 jrf 3 a
s5-8g &
S
1
1
TI
S 5
3 *j

1 sI'i
o> 8
1 fN (N

«"
i tj*
*• 3 S 2 8

1
"1 S -8 § =" OO
" 8
"8 o § «
y
B
1-s
2 S
V !_
> 3
OO
1 3*1 «s 2 (N 8
" a.
S

o
u

I |
1 a:
S ea ENR
67/33 ENR
55/45
a O 0 O Z
OA
454 MECHANICS OF PNEUMATIC TIRES
TABLE 6. 1 1 Comparison ofnatural rubber based tread compound with SBR based tread
compoundfor skid resistance on ice (from ref. [42])

Skidding on Skid on Cornering on


road ice lake ice lake ice
-2°C to -6°C -2°C -rc
OESBR 100 75SBR/25BR 100 100
OENR 131 OENR 117 121

6.5. Tire Wear


6.5.1 Theory of Tire Wear
As discussed in section 6.3.2, tire wear originates from sliding in the rear
part of the contact patch. The abrasion taking place there will be assumed
to be independent of the sliding velocity so that it depends only on the
normal pressure distribution and on the total sliding distance covered by a
tread element. Figure 6.92 shows the deflection y of the surface of a slip
ping tire as a function of the position x on the undeformed circumference.
The slope of the straight line in the adhesion region x < X is
tan/8-Av/K (6.70)
The curve in figure 6.92 also describes the change in deflection of a tire
element during its passage through the contact area. If the element ad
hered to the road throughout, the deflection would follow the dotted
straight line to the rear edge of the contact. The difference z between its
ordinate and the actual deflection is therefore the distance slid when the
abscissa x is reached. The amount of z is read off figure 6.92 as
ya tan /? - y (6.71)

(a)

-40 -20 0 20 -40 -20 20


TRACK TEMPERATURE , *C

FIGURE 6.91 The skid coefficient of (a) oil extended NR and SBR tread compounds and (b)
unextended NR and SBR tread compounds as function of the track temperature including wet
and icy conditions (from ref. (74J) • SBR O NR.
TIRE TRACTION AND WEAR 455

and consequently, with eqs. (6.71) (6.59)


dz = (Av/ V — dy/dx) dx. (6.72)
The first term in eq. (6.72) is the kinematic sliding due to the slip alone,
and the second term represents the retraction of the tire surface to its un-
deformed state.
The pressure dependence of abrasion will be assumed to follow the em
pirical power law of eq. (6.24), repeated here for convenience:
(P/P0)" (6.73)
where h is the abrasion depth per unit sliding distance. If the contact area
is rectangular with the width b, eqs. (6.72) and (6.73) give the following
expression for the wear per unit travelling distance [48, 76]:

W = (b/v) / (p//»0)"[Av - (dy/dx) V] dx (6.74)

where X is the coordinate x at which sliding sets in. When this equation is
applied to the toothed wheel model of the tire used in section 6.3, and the
pressure distribution is again assumed to be elliptical, the result is [77]
W = 1/2 ab (4L/7T abpa)" sF(c) (6.75)
where
p"c-_
for c < 1 F(c) = f(c) +
n + \ (1+
(6.75)
for c > 1 F(c) = /(c) + 1-P"
n+ I (1 + c2)"

and

/(O-O-etyO+cO/'o-. du
»0

and c has been defined by eqs. (6.34) and (6.35); ρ is the resilience of the

LATERAL
TIRE
DEFLECTION

POSITION ON CIRCUMFERENCE, x —•-

FIGURE 6.92 Deflection and sliding in the contact area of a slipping tire.
456 MECHANICS OF PNEUMATIC TIRES

whole tire. It appears here as an inverse measure of the dynamic losses in


the tire as it passes through the strain cycle imposed by the slip. Only part
of the elastic energy stored in the front part of the contact can be con
verted, in the rear part, into external work against the friction forces, the
rest being lost as heat.
The resilience effect can give a grossly misleading estimate of the abra
sion resistance of a tread material when tested on a machine using slipping
wheels as samples because the resilience of the tread material alone comes
then into operation [58]. The resilience of a tire, however, is dominated by
that of its pneumatic chamber, and is only partially affected by the tread.
At small slip and/or high load, eq. (6.75) reduces to

1
W-- (6.76)
n+ 1
For n = 1, eq. (6.76) gives a square law dependence of W on s; this re
sult was obtained in the first publication on this subject in which abrasion
was assumed to be proportional to the fractional energy dissipation [58].
Figure 6.93 shows how the function F(c) in eq. (6.75) depends on c at
various resiliences. The exponent n in eq. (6.73) had been put equal to 3/2
which is, according to table 6.4, a representative value. All curves in figure
6.93 approach asymptotically the same limit at large c when sliding ex-

0.02

FIGURE 6.93 F (c) in equation (6. 75) asfunction ofcfor different resiliences (from ref. [48]).
TIRE TRACTION AND WEAR 457

FIGURE 6.94 Theoretical slip- and load dependence of the wear of a car tire 7.50-15 (from
ref.[76]).
tends practically over the whole contact, and the wear becomes independ
ent of the resi Hence.
Eqs. (6.75) and (6.76) are valid for both lateral and longitudinal slip;
they will be evaluated here only for side slip. To carry out the calculations,
an assumption must be made about the load dependence of the contact
length a. On the basis of earlier work [48], the contact length has been
taken as proportional to the square root of the load. To avoid difficulties
with fractional powers of dimensioned quantities, an arbitrary reference
load L0 was introduced, and the contact length a0 under /,„ taken as
known [75]:
a = aa (L/Lo)">. (6.77)
With ρ = 1 and n = 3/2, eq. (6.75) gives the following dimensionless ex
pression for W
(ao6)l/2(Lo//>o)-V2 W - (W2) (L/A,)"4 sin 6F(c). (6.78)

6.5.2 Load and Slip Dependence of Wear


Slip enters eq. (6.75) through sinθ and F(c) so that, contrary to tire
forces, load and slip dependence cannot be jointly described by a univer
sal curve and recourse must be had to numerical examples. The grid dia
gram in figure 6.94 gives the theoretical load and slip dependence of the
7.50-16 passenger tire of figure 6.59 for which well documented c-values
were available. The reference load L0 was 453.6 kgf (= 1000 Ibs). Under
each load, wear initially increases rapidly with increasing slip, being pro
portional there to (P* according to eq. (6.76). The curves straighten at
larger slip but should asymptotically approach an upper limit for the rea
sons stated earlier.
The load dependence of wear varies greatly with the slip. It appears
strange at first that the load hardly affects wear at low slip. The reason is
that changes in load have two opposite effects; the contact length increases
with increasing load but the relative length of the sliding region (a - X)/a
458 MECHANICS OF PNEUMATIC TIRES

decreases at the same time. The load dependence of wear becomes pro
nounced, however, at large slip and finally approximates the power law of
eq. (6.73).
Slip dependence of wear cannot be determined in conventional road
tests and needs special means. A two-wheeled trailer was constructed
whose wheels could be set at slip angles in toe-in and toe-out positions
[77]. Results obtained with this equipment at slip angles up to 4° are given
in figure 6.95 on a doubly-logarithmic scale. Straight lines through the ex
perimental points have the slopes 2.4 for NR and 2.5 for SBR treads. Al
though these figures appear to agree well with theory, it must be remem
bered that the value of 3/2 for the exponent n is only an estimate, and that
the tire temperature increases with increasing sup because of the simulta
neous increase in power dissipation. This temperature rise could be partly
responsible for the increase in abrasion and will be discussed in section
6.5.4.
The interaction between load and slip dependence in figure 6.94 has not
be tested with tires but experiments were made with small solid wheels on
the same tracks on which the load dependence of ordinary sliding abra
sion had been determined (section 6.2.3). The load dependence of the
wear of these test pieces at any slip angle could also be described by a
power law
W - const. L" (6.79)
The exponents α shown in table 6.4 are always smaller than n and increase
with increasing slip angle (ordinary sliding corresponds to a slip angle of
90°). The theory is thus at least qualitatively confirmed.

04

FIGURE 6.95 Wear rate ofa car tire as aJunction ofthe slip angle on a trailer (from ref. [77]).
TIRE TRACTION AND WEAR 459

Practical interest is centered not so much in the slip dependence of wear


as in its dependence on the cornering power which the tire must develop
to keep the vehicle on a given curve of radius R at a velocity v, the equilib
rium being given by
v2/R = g (S/L) cos 0 (6.80)
where g is the gravitational acceleration. As the relation between ν2/R and
W cannot be written down explicitly, figure 6.96 gives numerical results
calculated again for the 7.50-16 tire of figure 6.94. The graph demonstrates
the great increase in wear with increasing radial acceleration and closely
resembles results by Cough and Shearer on a skid pad [78]. The slope of
the curves shows that wear can increase with the 6th power of the vehicle
speed. Results obtained under what may be described as severe service
conditions are reproduced in figure 6.97. They were determined on the
road track at Montlhery and show directly the strong speed dependence of
wear, which in this case increased with nearly the 4th power of the vehicle
velocity. It should be stressed that these speed effects are primainly due to
mechanical causes, high accelerations demanding large slips.

6.5. Effect of Tire Construction


If the wear of different tires is to be compared, their lateral stiffness
must be considered. The slip necessary for a given side force decreases

LATERAL ACCELERATION, VVR, % g

FIGURE 6.96 Theoretical dependence of wear on the lateral acceleration during cornering
under different loads (from ref. [76J).
460 MECHANICS OF PNEUMATIC TIRES

to

I"
bJ
(E

1*
40 50 60 70 80 90
VELOCITY, km/hr

FIGURE 6.97 Relative tire wear asJunction of vehicle speed on a race track (from ref. 158]).

with increasing stiffness ks; eq. (6.36) shows it to be inversely proportional


to ks at small slip. Eliminating the slip between eqs. (6.36) and (6.76) the
following relation between wear, side force and stiffness is obtained:

W-- ab(abp0Y (6.81)


n+ 1

which holds only for S <sc fiL. If, therefore, two tires 1 and 2 are tested at
equal slip, and they differ in no other respect than their stiffness kn their
wear rates are in the ratio [(ks)1 /(&,)J", from eq. (6.76). If they are com
pared at equal side force, the ratio is (ks)2/(ks)1 according to eq. (6.81); in
other words, their rating is reversed as compared with the test at equal
slip.
The difference between results of equal slip and equal side force can be
demonstrated with the trailer previously referred to. When two nominally
identical tires are mounted on it, they obviously run at equal slip and
equal side force. When the tires differ in their lateral stiffness, the trailer
automatically orients itself to run at equal side force [58], [77]. Figure 6.98
gives results obtained with diagonal and radial tires, both with NR treads.
The initial cornering stiffness 1/2 k/f of the radial tire was about
greater than that of the diagonal tire.
TIRE TRACTION AND WEAR 461

The dashed line curves give the wear at equal slip. The radial tire hap
pened to be better even under this condition but the difference between
radial and diagonal tires is greatly increased at equal side force (solid
lines), that is, under service conditions.
There is a limit to useful lateral tire stiffness, in spite of its advantages in
wear, because a state is eventually reached in which steering becomes pre
carious. The reasons have been indicated in section 6.3.1, and are also dis
cussed in ref. [71].
6.5.4 The Relative Wear Rating of Tread Compounds
The life of a tire depends on a wide range of factors, many outside the
control of the experimenter. Evaluation of the wear resistance of tread com
pounds is, therefore, invariably on a comparative basis. The tread height
loss of tires of the same construction but with different tread compounds is
determined under the same experimental conditions and the distance cov
ered per unit tread height loss is calculated. The index most commonly
used is the relative wear rating, defined as the ratio between the wear of
the control and that of the experimental compound.
The relative wear rating of two tread compounds depends pronouncedly
on the testing conditions such as cornering and traction forces, topology of
the track surface, ambient temperature and wetness of the road surface.

so

Equol Slip
Equal Side Fores

40

30

nT 20

10

I 2 3
"SET" SUP ANGLE , degrees

FIGURE 6.98 Wear rate of bias and radial ply tire, both with NR treads when running either
at equal slip or at equal side force (from ref. [77]).
462 MECHANICS OF PNEUMATIC TIRES

Their interaction is so complex that for practical purposes it is still neces


sary to carry out the test under conditions which should closely match
those met in practice.
Even so, the research into the nature of abrasion and into the mecha
nism of tire wear, discussed in section 6.2 and 6.5, allow a considerable ra
tionalisation.
Figure 6.94 showed that up to 4° slip angle there is virtually no load de
pendence of tire wear. Most steering, overtaking manoeuvres, braking and
accelerating manoeuvres take place at accelerations under 0.2 g. This cor
responds with radial tires to slip angles below 2° and slip values for longi
tudinal slip of less than 5% (see figure 6.85). It would, therefore, be ex
pected that the rate of wear does not critically depend on the tire load
except that a heavier vehicle needs larger forces on the tire to achieve the
same acceleration. Interaction between compounds and tire load is very
unlikely and therefore equation (6.76) can be taken to describe the wear
rate of a compound on a tire with sufficient accuracy.
The relative wear rating becomes then
W
-57 = 2 exp(n, - n2) (6.82)
n, + 1

where suffix 1 refers to the control and suffix 2 to the experimental com
pound. (For the same tire, a, b and k are the same). Also, the resilience of
the compound in equation 6.76 is to the most part determined by the resil
ience of the tire so that it is assumed here to be equal for the both tires.
The expression /t /;,, can be taken as a measure of the energy necessary
to abrade unit volume of rubber per unit sliding distance. A severity de
pendence of the relative wear rating is indicated because of its dependence
on the slip. For a comparison between NR and SBR, this means, accord
ing to table 6.4, that SBR becomes progressively worse in relation to NR
with increasing slip angle and this should be even more pronounced in a
comparison between BR and NR. In figure 6.95, wear data of NR and
SBR vs. slip angle were shown on a double logarithmic scale. The two
straight lines are running virtually parallel, indicating an independence of
the relative wear rating on slip angle or perhaps there is a very small in
dication that the relative wear rating of SBR vs. NR should decrease with
increasing slip. Figure 6.99 shows a further comparison of NR with SBR
with the same trailer as in figure 6.96, but obtained during a period of
much lower ambient temperature. Included are also data for a 50 NR-/50
BR blend; a tire tread compound made wholly of BR is not feasible,
mostly because BR has a poor tear resistance, leading to chipping and rib
tearing. Both SBR and the blend are better than NR at high slip angles
but are poorer in wear resistance at lower slip angles; although the power
indices in table 6.4 would suggest the reverse. This is also borne out in
practice. NR/BR blends show an improved wear resistance over NR only
in severe service and/or at high ambient temperatures.
Another effect appears therefore to intervene in the slip angle depen
dence of tire wear. In fact, the intrinsic wear resistance (μp0)n of the tread
compound depends on temperature, as demonstrated in figure 6.100 which
shows the volume loss per unit distance of an SBR and a natural rubber
compound as function of the ambient temperature. The experiments were
also carried out with the trailer but at a constant slip angle θ, constant
speed and on the same type of road surface, but during a dry period at
TIRE TRACTION AND WEAR 463

FIGURE 6.99 Wear rate as function of the slip angle on a trailer for three different tread
compounds at low ambient temperatures (from ref. [79]).

which the ambient temperature rose steadily. The wear of both com
pounds increased as the temperature increased, more so for the natural
rubber than for the SBR tread stock. From the straight line graphs, tem
perature coefficients can be derived which are about 3% per °C for the NR
but only 1.5% per °C for the SBR stock. Such temperature dependence of
wear of tread compounds has been reported in the literature [80, 81, 82]. It
can lead to complete reversal in the ranking of the wear rating, as is in
deed apparent from the above figure, and it has therefore been proposed
that, for best wear results, different rubbers should be used in different
geographical regions: Natural rubber—based compounds being employed
essentially in the colder regions and SBR or SBR/BR blends in the
warmer parts of the world. Indirect evidence suggests that BR should have
only a small abrasion temperature coefficient and one would therefore ex
pect that it would reduce the temperature dependence of natural rubber or
even of SBR if blended with it.
The temperature of the tire, apart from depending on ambient and road
surface temperature, rises above their level because of internal heat build
up and factional energy dissipation in the contact. What matters for wear
is the resulting tire surface temperature which takes some time to become
established and which generally differs greatly from ambient and road
surface temperature.
Figure 6.101 shows the slip angle dependence of the equilibrium tem
perature rise of a natural rubber and a SBR tire after the tires had been
run for a sufficiently long time. The temperature was determined by stop
ping the trailer [77] and placing a temperature probe on the surface. The
linear increase in temperature with slip angle has been explained by
Schallamach [83], who also attempted to relate this measurement to the
actual temperature as it occurs during sliding in the contact region. He
finds that these two temperatures should be proportional to each other, al
though the latter is not amenable to direct measurement. It reaches a very
high values in the sliding region of the contact area, cooling taking place
464 MECHANICS OF PNEUMATIC TIRES

60 70 80
TIRE SURFACE TEMPERATURE, «C

FIGURE 6. 100 Wear of an NR and SBR tread compound on a trailer at a slip angle of 1 ° as
Junction of the ambient temperature (from ref. (77]).

2 3
SLIP ANGLE, dag.

FIGURE 6.101 Tire surface temperature of different tread compounds as function of the slip
angle (from ref. [77]).
A - SBR, H - NR, C - JO NR/50 BR
TIRE TRACTION AND WEAR 465

1 1 1

•— • 1 CONTACT
1 -ENGTH
1.5
E

/\
V
iO.5
5 /
CO
0 1_J** J
u'V-W
-O5 1 1 1 1 1
2 4 6 8 10 12 14 16 18 20
DISTANCE TRAVELLED, cm

FIGURE 6. 102 Sliding path of a tire element in the contact at a slip angle of 1 ".

after leaving the contact area, and in the front part of the contact area,
when tire and road are in adhesive contact.
Figure 6.102 shows the deflection relative to the track of a point on a
tire running at a slip angle of 1°.
Near the rear end of the contact sliding sets in, the ordinate of the graph
giving the sliding path. From the travelling speed of the tire and the slid
ing path, it is possible to calculate the sliding speed of this particular point
on the tire while in the contact; and from temperature vs. sliding speed
measurements it is possible to estimate the temperature rise which this
sliding will cause. Only an estimate is possible because it is ; transient
phenomenon and the actual temperature rise will depend on the thermal
conductivity of tire and track. From figure 6.102 it can be deduced that the
sliding speed is
K, = 0.08 V forward
ie., at a forward speed of 50 km/h the sliding speed would be about 4 km/
h or the expected temperature in the contact area could reach 100° C (de
duced from figure 6.42).
Assuming that the wear of slipping wheels follows the original equation
of Schallamach and Turner in which the severity effect due to the non-lin
ear load dependence was neglected, but allowing for the temperature rise
due to an increase in slip angle by putting
W=K9i(\ (6.83)
in which Δt is the temperature rise, the temperature coefficient and the
constant K can be calculated.
Table 6.10 shows these constants for the natural rubber, SBR and a nat
ural rubber/BR blend [77]. It is immediately obvious that they are very
close in magnitude to those found independently by the direct measure
ments shown in figure 6.100.
Figure 6.103 shows the wear of tires with NR and SBR tread com
pounds on the same road surface when dry and wet under otherwise con
stant conditions of load and slip angle and at almost the same ambient
temperature.
466 MECHANICS OF PNEUMATIC TIRES

30

50 100
% WET ROAD

FIGURE 6.103 Wear rale of NR and SBR tread compounds at various percentages of road
wetness under otherwise equal conditions (from ref. [78]).

In the range between these two extreme conditions, stretches of wet and
dry road and partly wet road were met with to varying degrees. To quan
tify these conditions, the % wetness was defined as the ratio of wet road to
total distance covered. It is seen that on the dry road, SBR compound had
the higher wear resistance whilst the reverse was true on the wet road.
Similar reversals have been reported by Westlinning [84]. Several ex
planations of this phenomenon are possible. The authors proposed that
the tire surface temperature was drastically reduced by the water on the
road [77] and although the wetness affected the absolute rate of wear, the
reversal in rating was essentially due to the different temperature coeffi
cients of wear of these two rubbers. Other investigators, however, see in
this behaviour a shift in the importance of two different abrasion mecha
nisms [85], fatigue and abrasive wear, and it must be admitted that this re
mains a possibility, although it is not open to direct experimental veri
fication.
Indeed, if the relative wear rating of NR vs. SBR obtained in trailer
measurements over a wide range of slip angles and weather conditions is
plotted as function of the tire surface temperature, a single curve results,
independent of the testing conditions, as seen in figure 6.104, curve (A).
The graph also includes results of conventional road tests (open circles)
where the slip values are no longer under the control of the experimenter
TIRE TRACTION AND WEAR 467

[86]. These findings also hold for other compound comparisons as seen
from curves (B) and (C), which were obtained under different weather (ie.
wet and dry road) and driving conditions. This is particularly impressive
for the 2 points marked (a) and (b) in figure 6.104, both of which have
about the same wear rating of OENR relative to OESBR. The rating (a)
was obtained with the trailer at a constant slip angle of 2° and a speed of
30 mph, whilst point (b) was obtained in a car test on a mainly straight
motorway. Although the absolute loss rate differed by more than a factor
of 20, the rating was nearly the same in both cases, the only common pa
rameter being the same tire surface temperature.
Undoubtedly, there are severity effects which influence the wear rating of
tread compounds, but it appears that temperature plays such an important
role that it can reverse the ranking of compounds in a way which could
not have been attributed to severity effects such as surface and pressure
dependence of the abrasion rate.

6.5.4 Unequal Tire Wear


In the previous discussions, the tire was taken to be a cylinder with a
uniform pressure in the axial direction and—in the theory of Schallamach
and Turner [58]—an elliptical pressure distribution in the circumferential
direction.
In reality, the tire has also a curvature in the axial direction which
creates a pressure distribution across the tire in the contact area. The exis
tence of a tread pattern results in local pressure patterns: At the edge of a
rib or block, the pressure drops to zero.

30 40 50 60 70
TIRE SURFACE TEMPERATURE/'C

FIGURE 6.104 The relative wear rating of tread compounds as Junction of the lire surface
temperature; (A) NR vs. SBR; (B) OENR vs. OESBR and (C) OENR/BR vs. OESBR/BR.
• • rraulLs obtained with a constant slip angle trailer under high Kverity.
O—O tot car result! (from ref. |86|).
468 MECHANICS OF PNEUMATIC TIRES

The double curvature at the crown of a tire creates shear forces between
the tire and the contact area both in the axial and the circumferential di
rection as demonstrated in figure 6.105 which shows the pressure distribu
tion and the axial and circumferential shear stresses between tire and road
at several points in the contact of a stationary tire. Such diagrams are well
known and for a more detailed discussion the authors refer to the litera
ture [88, 89]. The shear stresses are modified by the tread pattern; because
rubber is virtually incompressible but has a low Youngs modulus, a rub
ber block bulges out under load. This produces outward-directed shear
stresses at the interfaces between track and rubber and the frequent
changes in direction of the lateral stresses in figure 6.105 bear witness to
this effect. Because the pressure tends to zero at the edges of the block,
there is a tendency for sliding even in free rolling tires resulting in the typ
ical wear patterns of heavy service tires on front wheels or trailer axles
along the block or rib edges as shown in figure 6.106.
The contact forces are considerably modified when the tire has to trans
mit lateral forces, as shown in figure 6.107. The three columns show the
lateral contact forces along the contact area at different points across the
tire in free rolling and at 1° and 2° slip angle. If the lateral force is small at
0° slip, the stress at 1 ° slip rises linearly from the front toward the exit of
the contact, as would be expected from the theory outlined in section 6.5.1.
If the tire already has a shear force at zero degree slip angle due to the

Pressure Lateral Shear Longitudinal


Stress Shear Stress

n v

\S

\
n
LA
1 \S
Stress distribution in the contact area of a 10.00 R
20 tire, Load: 2830 Kgf.,Pressure: 5.8 bar,
0° Slip angle.

FIGURE 6.105 Preaun, lateral and circumferential stress al various points in the contact of a
patterned tire.
TIRE TRACTION AND WEAR 469

FIGURE 6. 106 Wear pattern due to sliding at the edges of the tread pattern of a heavy service
tire, run only on front axles.

crown radius of the tire or because of block squeezing, the resultant shear
force at a given slip angle is the superposition of the shear force at zero
degree and the stress due to imposed slip. The sliding path which then re
sults from a slip angle is shown in figure 6.108.
There are even some small sliding movements with no slip angle, nota
bly at the edge of blocks as explained above. Even if the tire rolls at a slip
angle, there is generally adherence in the front of the contact area and a
rapid increase in the sliding path towards the end of the contact, as al
ready discussed in the theory of slipping wheels in section 6.5.1. This is
true for almost all measured points when the tire runs at one degree slip
angle. Sliding at 0° slip angle appears to be superimposed on the sliding at
one degree slip angle. If the sliding path at zero degree is opposite to that
produced by the slip angle (for example measuring point 3), then sliding is
noticable at 1 ° slip angle in the front of the contact area in the same direc
tion as for 0° slip angle. Adhesion takes place in the center section and
sliding in the direction of the slip angle sets in very abruptly at the rear of
the contact area. If sliding takes place at zero degree slip angle in the di
rection of that produced by the slip angle, sliding sets in at the front of the
contact area and continues over the whole length at an almost uniform
rate (measuring point 2). The maximum recorded sliding path is shorter
than would have been expected from simple geometry. For a contact
length of about 130 mm, a maximum sliding path length of 2.3 mm would
470 MECHANICS OF PNEUMATIC TIRES

Path through
contact area

FIGURE 6.107 The lateral stress at 0°, 1 ° and 2° slip angle in the contact at various points of
the lire tread pattern.

have been expected. In fact, it never exceeded 1.5 mm. At 2° slip angle,
the sliding paths are longer, as expected, and sliding sets in earlier in most
cases; when sliding at 0° has the same direction as at slip angles, the re
sulting sliding extends over practically the whole contact area. The maxi
mum possible values of 4.5 mm are more nearly reached than at one de
gree slip angle.
The abrasion at a point on the tire is function of the energy dissipated in
sliding. Eq. (6.24) states that the volume loss per unit area and sliding path
distance is proportional to the nth power of the pressure.
Investigations on the pressure dependence of the friction coefficient
have shown that it decreases with the 1/3 power (eq. 6.3) on smooth, and
on rough surfaces with the 1/9 power of the pressure. The abrasion loss
would, therefore, vary with the energy dissipation E as

(6.84)
(Mo/Jo)"

where fi0p0 is a material constant (see also section 6.5.1, eq. 6.73).
If the small pressure dependence of the friction coefficient is neglected,
eq. 6.84 becomes

(6.85)
TIRE TRACTION AND WEAR 471

0* . 1" . 4-

100 mm

FIGURE 6.108 Lateral sliding at 0°, 1° and 4° slip angle in the contact patch at various points
of the tire tread pattern.

From measurements of the shear forces H and the sliding path S as a


point passes through the contact area, the energy dissipation can be calcu
lated as

£= 1° fi-ds (6.86)
»0

or, if only the pressure and the sliding path were known,

(6.87)

where Sx and Sy are the sliding path components in the lateral and circum
ferential direction respectively. In general, interest on wear behaviour cen
ters on the problem of different rates of wear between different sections of
tread pattern, notably between center and shoulder regions. The material
constant μop0 is the same for both points and the relative rate of wear be
tween these two points is given by
472 MECHANICS OF PNEUMATIC TIRES

It will be noticed that the material constant n tends to exaggerate any


tendency to uneven wear.
Figure 6.109 (a) and (b) shows pressure measurements and lateral slip
path measurements, both at two degree slip angle, for two points (loca
tions 1 and 6 of figure 6.108), along the direction of travel of the tire. From
these measurements the integral in equation (87) was evaluated numeri
cally as

E = £ Pi AS, (6.88)
i-O

where the index / refers to the ith equally spaced interval along the circum
ferential direction of the contact path, pi · Δ Si is shown in figure 109(c)
as function of the circumferential contact path. The area under the curve
is then the integral of equation (6.87). Only the lateral slip path Sy was
used because Sx was always much smaller.
Evaluated according to equation (6.88) the area for the point at the cen
ter line of the tire is 19.6 and for the point in the shoulder region is 26.2
units. Under a slip angle of 2°, the shoulder of this tire would consume
34% more energy in sliding than the center, giving rise to about 55% more
wear, if n is taken to be 1.5 as in section 6.2.3.
Of course, a tire runs under a variety of slip conditions so that the above
is an extreme estimate. An average should be taken over several slip an
gles and several measuring points, and with the slip angle in both direc-

Shoulder Region Center Region

CONTACT PATH •

FIGURE 6. 109 Pressure, sliding path and p • A S at points in shoulder and center region of a
tire contact.
TIRE TRACTION AND WEAR 473

t ions to allow a more realistic estimate of the likely magnitude of uneven


wear to be encountered in practice.
Since uneven wear is much more a problem on free rolling axles ie.
front- and trailer wheels, a measurement under circumferential forces and
hence slip is not essential but is of course possible; the method is the same.

References

1 P. Thirion, Rev. Gen. Caout. 23 (1946), 101


2 A. Schallamach, Proc. Phys. Soc. B 65 (1952), 657
3 E. Southern and R. W. Walker, Pol. Sci. Technol. 5 A (1974) 223
4 F. P. Bowden and D. Tabor, "Friction and Lubrication of Solids". Oxford Univ. Press
(London, 1952)
[5] A. Schallamach, Wear 1 (1958), 384
[6] A. Schallmach, "Abrasion and Tire Wear" in L. Bateman, Ed. "The Chemistry and
Physics of Rubberlike Substances" McLaren & Sons, LTD., London (1963), 355
[7] W. Gnörich and K. A. Grosch in the press
[8] A. Schallamach, Wear 13 (1969), 13
[9) A. Schallamach, Wear 17 (1971), 301
[ 10] A. Schallamach, Gummi- Asbest-Kunst. 28 (1975), 142
[1 1] K. A. Grosch: Proceedings of the Royal Society A 274 (1963), 21
[12] M. L. Williams, R. F. Landel and J. D. Ferry, Journ. of American Chem. Soc., 77
(1955), 3701
[13] J. D. Ferry: Visco-elastic Properties of Polymers J. Wiley & Sons, Inc. (1961)
[14] T. L. Smith: Journ. of Pol. Sci., 32 (1958), 99
[15] L. Mullins: Trans IRI, 35, no. 4 (1959), 213 (1965)
[16] K. A. Grosch and A. Schallamach: Trans. IRI 41 (1965) T 80 also Rub. Chem. and
Techn. 39, (1966), 287
17] K. A. Grosch: Ph.D. Thesis, Univ. of London (1963)
18] W. Gnörich and K. A. Grosch: Journ. of the Institute of the Rubber Industry, 6, (1972X
192
19] F. A. Greenwood and D. Tabor: Proc. of Phys., Soc. 71 (1958), 989
20] A. Schallamach, Rubber Chem. Techn. 41 (1968), 209
21] D. F. Moore and W. Geyer, Wear 22 (1972), 1 13
22] A. Schallamach, Wear 6 (1963), 375
23] M. Barquins, D. Maugis and R. Bourtel C. R. Acad. Sci. 279 (1974) Serie B, 565
24] M. Barquins and R. Bourtel, Wear 32 (1975), 133
25] A. D. Roberts and A. G. Thomas, Wear 33 (1975), 45
26] G. A. D. Briggs and B. J. Briscoe, Wear 35 (1975), 357
27] A. D. Roberts and S. A. Jackson, Nature 257 (1975), 118
28 G. A. D. Briggs and B. J. Briscoe, Nature 262 (1976), 381
29' A. D. Roberts and A. G. Thomas, NR Techn. 7 (1976), 38
30 A. Schallamach, Wear 35 (1975), 375
31 J. A. C. Harwood, L. Mullins and A. R. Payne T. Instr. Rubber Ind. 1 (1967), 17
32] A. Schallamach, Proc. Phys. Soc. B 66 (1953), 817
33] A. R. Savkoor and T. I. Ruyter A. C. S. International Symposium on Polymer Friction
and Wear, Los Angeles (1976)
[34] K. A. Grosch: The speed and temperature dependence of rubber friction and its bearing
on the skid resistance of tires in: The Physics of Tire Traction (Hays, D. F. and
Browne A. L., Eds.) Plenum Press (New York-London 1974), 143
[35] A. D. Roberts: Lubrication studies of smooth rubber contacts. The Physics of Tire Trac
tion, Theory and Experiment, ed. by D. F. Hayes and A. L. Browne, Plenum Press
(New York-London 1974), 143
[36] A. G. M. Mitchell "Viscosity and lubrication. The Mechanical Properties of Fluids, ed.,
Brackie & Sons LTD., London 1944
M. T. Kirk and T. F. Archad, Proc. Roy. Soc., London 261, (1961), 532
A. Cameron, "The Principles of Lubrication, Longman, London
W. Geyer, Automobil Industrie 17, 2 (1972), 41; 17, 4 (1972), 39
E. Southern, and A. W. Walker: Natural and Physical Science 237, no 78, (1972), 142
H. Rieger: Experirnentale und theoretische Unter-suchungen zur Gummireibung, Dis
sertation, TH Munich (1968)
474 MECHANICS OF PNEUMATIC TIRES

[42] K A. Grosch: Oil extended Natural rubber in Winter tires:—theory and practice, Int.
Rubber Conference San Francisco, 1976
[43 A. SchaUamach, Rubber Chem. Techn. 41 (1968), 209
[44 A. SchaUamach, Trans. Inst. Rubber Ind. 28 (1952), 256
45 A. SchaUamach, Jour. Pol. Sci. 9 (1952), 385
[46 A. G. Thomas, Journ. Polym. Sci. 18 (1955), 177
[47 I. V. Kragelskii and E. F. Neponnyashchii, Wear 8, (1965), 303
[48] K. A. Grosch and A. SchaUamach: Kautschuk, Gummi und Kunstst. 22 (1969), 288
[49] G. J. Brodskii, Sakhuoskii, M. M. Reznikovskii and V. F. Evstratpv, Soviet Rubber
Technol. 19 (1960) B, 22
50] A. SchaUamach: J. Appl. Polym. Sci. 12 (1968), 281
51 A. N. Gent: Journ. Appl. Polym. Sci. 6 (1962), 497
52 G. J. Lake and A. G. Thomas: Kautschuk Gummi Kunstst. 20 (1967), 211
53 E. Southern, Private Communication
54 D. H. Champ, E. Southern and A. G. Thomas, Coatings and Plastics Prep. 34, NOI
(1974), 234
[55] E. Southern and A. G. Thomas, Leeds- Lyon Symposium on Tribology" The Wear of
Non-Metallic Materials" (1976)
[56] R. S. Rivlin and A. G. Thomas, Journ. of Pol. Sci 10 (1953), 291
A. SchaUamach, Rubber Chem Tech. 33 (1960), 854
A. SchaUamach and D. M. Turner, Wear 3 (1960), 1
V. E. Gough, Auto Eng. 44 (1954), 137
B. J. Allbert and J. C. Walker, Proc. IME 180 (1965-66), 105
61] V. E. Gough and T. French, Proc. First Inter. Skid Prevention Conf. CharlottesviUe, Va
(1959) Part I, 189
62] A. SchaUamach, Rubber Chem. Techn. 43 (1970), 995
D. L. Nordeen and A. D. Cortese, Trans. S.A.E. 72 (1964), 325
A. SchaUamach, unpublished work
65] H. B. Pacejka, this book, chapter 8
[66] D. J. Schuring, Tire Sci. & Techn. 4 (1976), 115
[67] N. Kienle "The Role of tread pattern—A blend of the simple and the complex" in "The
Physics of Tire Tract. Theory and Experiment" D. F. Hayes and A. L. Browne, Eds.
Plenum Press, New York 1974
[68] W. Gengenbach, Automobil-technische Z. 3 83; 8 288; 9311 (1968)
[69 G. Maycock, Proc. Inst. Mech. Eng. 180 (2A) (1965-66), 122 Rubber Chem. Techn. 41,
(1968), 780
K. A. Grosch and A. SchaUamach, Rubber Chem & Techn. 49, (1976), 862
71] C. G. Giles, Highway Res. Rec. 46, (1964), 43
72] B. Förster, Deutsch. Kraft. Forschung, Zeischenbericht 22 (1938)
73] T. K. Meades, Brit. Road. Res. Lab. Report LR 73 (1967)
74] K. A. Grosch and G. Maycock: Trans. Inst. Rubber Ind. 43 (1966), 280, Rubber Chem.
& Techn. 41 (1968), 477
[75] R. F. Peterson, C. F. Eckert and C. I. Carr "Tread Compound Effects in Tire Traction"
in "The Physics of Tire Traction, Theory and Experiment" D. F. Hayes and A. L.
Browne, Eds. Plenum Press, New York 1974
[76] A. SchaUamach, "Mécanique Materiaux Electri. (1972) (265-1) Numero spécial l'usure,
77 Gummi Asbest Kunstst. 25 (1972), 442
K. A. Grosch and A. SchaUamach, Wear 4 (1961), 356 (Tire wear of controlled slip)
[7 V. E. Gough and G. R. Shearer, Inst. Mech. Eng., Proc. Autom. Div. 6(1955-56), 171
T78
79 K. A. Grosch: in the press
[80 J. Mandel et al Ind. Eng. Chem. 43 (1951), 2901
[81 H. C. J. deDecker et al. Proc. of the 3rd Int. Rubber Techn. Conference, London 1954,
749
[82] C. Prat, Re. Gén. Caoutch. 32 (1955), 991
[83] A. SchaUamach, Inst. Rubber Ind. 1 (1967)
[84] H. Westlinning, Kaut. Gummi WT (1956), 273
[85] D. Bulgin and M. H. Walters, Fifth Int. Rubber Conference, Brighton 1967
[86] K. A. Grosch, J. Inst. Rubber Ind. 1 (1967), 35
[ 87] K. A. Grosch "The assessment of wear and skid resistance of tire tread compound" in
Development with Natural Rubber ed. J. A. Brydson McLaren & Sons, London
(1967)
[88] M. Gerresheim, Automobil Industrie 20 (1975), 59
[89] N. Seitz and A. W. Hussmann, SAE paper no 710626 (1971)
[90] B. Förster, Deutsch. Kruftf. Forsch. 22 (1938)
Chapter 7
TIRE STRESS AND DEFORMATION
R. A. Ridha1
S. K. Clark2

7. 1 Introduction 477
7.2 Tire Construction 477
7.3 Tire Loading 479
7.4 Tire Response 480
7.5 Analytical Models 480
7.5. 1 Early Tire Models ' 48 1
7.5.2 Netting Analysis 482
7.5.3 Limitations of Netting Analysis 488
7.5.4 Membrane Analysis 489
7.5.5 Limitations of Membrane Analysis 490
7.5.6 Thin Shell Analysis 490
7.5.7 Limitations of Thin Shell Analysis 494
7.5.8 Finite Element Models 496
7.5.8. 1 Constant Strain Triangular Element 499
7.5.8.2 Element Stiffness Matrix 500
7.5.8.3 Solution 502
7.5.8.4 Applications 503
7.5.8.5 Nonaxisymmetric Analysis 504
7.5.8.6 Applications 506
7.5.8.7 Limitations of Constant Strain Models 506
7.5.8.8 Shell Elements 508
7.5.8.9 Limitations of Shell Elements 509
7.5.8.10 Isoparametric Elements 511
7.5.8.11 Element Description 513
7.5.8.12 Displacement Field 513
7.5.8.13 Two-Dimensional Analysis 515
7.5.8. 14 Variable Material Properties 515

1 Formerly with Firestone Tire and Rubber Co., Central Research Laboratories, Akron, Ohio. Now it Retearch Labora
tories. General Tire and Rubber Co.. Akron, Ohio 443 17.
2 Dep'l.. Mechanical Engineering, University of Michigan, Ann Arbor, Michigan 48109

«7S
476 MECHANICS OF PNEUMATIC TIRES

7.5.8. 1 5 Applications 516


7.5.8.16 Tire Inflation 516
7.5.8.17 Load-Deflection 519
7.5.8.18 Future Developments in Analysis 522
7.6 Experimental Techniques 522
7.6. 1 Introduction 522
7.6.2 Measurement of Surface Strain 523
7.6.3 Measurement of Internal Strain 527
7.6.4 Measurement of Cord Loads 529
7.6.5 Measurement of Bead Forces 532
7.7 Perspective 533
7.8 Notation 536
7.9 Acknowledgements 537
References 537
TIRE STRESS AND DEFORMATION 477

7.1 Introduction
The pneumatic tire is often taken for granted as a simple and reliable
component of the vehicle. A closer look, however, shows that the tire in
service is subjected to severe stresses and deformations whose quantities
must be determined in order to accurately predict tire performance.
Modern tire structures have evolved through a series of modifications to
the original rubber tire. These modifications were based on field experi
ences and on mostly experimental studies of tire behavior. Use of analyti
cal techniques to calculate tire stresses and deformations remained limited
in scope for a long time because the complexity of the tire structure placed
it beyond the domain of available methods of analysis. The recent gain in
emphasis on analytical techniques is due, at least partly, to their potential
for becoming less time consuming and less expensive than experimental
methods, the need for predicting a tire's behavior before its manufacture,
and the notable advances in computational and structural analysis meth
ods. In this chapter, these methods are described and their applications
shown for calculating tire stresses and deformations.
Structural analysis is the analytical determination of structural re
sponses to a prescribed set of applied loads. The responses may be dis
placements or distortions if force loads are known, or forces if dis
placement or distortions are known. Given the geometry of a structure
(shape, dimensions), the relevant properties of its component materials,
the magnitude and distribution of applied loads, and any constraints from
boundary conditions, then structural analysis is used to calculate dis
placements, strains, or stresses at any chosen location in or on the struc
ture. These calculated values may be compared to those required for func
tionality of the structure. Although structural analysis is not directly
applicable to determining the most efficient configuration of the structural
components, the analysis of successive well-chosen modifications can of
ten optimize compositions or geometries.
Application of structural analysis to a tire requires (a) knowledge of the
relevant physical properties of the component materials, and their config
uration in the tire, ( b) complete characterization of the applied loads, and
(c) an analytical technique (i.e. theory) for calculating the required re
sponses. These requirements are explained in the following sections.

7.2 Tire Construction


One arrangement of the principal components of a tire is shown in Fig
ure 7.1. Each ply is a layer of parallel cords embedded in rubber. When
the tire is vulcanized under pressure it becomes a unit through which
stresses and strains can be transmitted. In addition to the ply rubber, a tire
has tread and sidewall rubbers, in the locations shown by dotted lines in
Figure 7.1. These usually have compositions different from each other.
Still different compounds are used in the innerliner, bead filler and inserts.
In addition to material variations, the tire engineer has many options
available for modifying tire properties. Perhaps the most powerful is that
of changing cord angles in either the body or tread plies. This has led to
three main types of tires: bias, belted bias, and radial.
As shown in Figure 7.2, the body cords in a bias tire make a rather large
angle with the tread centerline, and there are no tread plies. Bias tires for
471 MECHANICS OF PNEUMATIC TIRES
TREAD STOCK

TREAD (BELT)
PLIES
SIDEWALL
STOCK

BEAD FILLER
RIM

FIGURE 7.1 Cross-section of a belled tire

passenger cars usually have either two or four plies with cord directions
essentially symmetric to the centerline. Each ply is usually anchored
around the bead.
Belted bias tires (Figure 7.3) have the basic two-ply or four-ply body of
bias tires plus two or more belt plies between the tread rubber and the tire
body. Cords in the belt plies are more nearly circumferential than those in
the body plies. The belt constrains both the width and circumference of
the tread area, thereby restricting the inflated tire profile and reducing
tread movement in the footprint. Tread plies at angles comparable to
those of body cords are often called breakers rather than belts. In either
case they improve impact resistance.
Radial tires (Figure 7.4) are also belted, but have body cords that lie en-

FlGURE 7.2 Cord pattern in a bias tire


TIRE STRESS AND DEFORMATION 479

FIGURE 7.3 Cord pattern in a belted bias tire

tirely in meridional planes of the tire; i.e., perpendicular to the tread cen-
terline. This results in an extremely flexible sidewall which acts independ
ently of the belt, thus reducing tread movement in the footprint to less ^ *c^
than that in belted-bias tires. /H. iw^v^n-Mr n- ^juA • sOt**
Tires are being developed which contain no fabric, only an elastomer to
form the air chamber and bead wires for anchoring to the wheels [I]. De
formation and performance characteristics are altered by varying the
thickness distribution in the body, which is somewhat analogous to vary
ing cord angles in conventional tires.

7.3 Tire Loading


Loads applied to pneumatic tires include inflation pressure, externally
applied mechanical loads, and thermal loads.
Inflation of a tire subjects its inner surface to a uniform outward pres
sure that accounts for most of the total load on a tire in service. Dynamic
loads are superimposed on this static load.
Mechanical loads include static and dynamic forces applied to the tire
at the tire-road and tire-rim interfaces, and forces caused by rotation, im
pact, and vibration of the tire. These loads are thus the actions and reac
tions resulting from the tire surface in contact with the road and wheel
rim. Load imposed by the vehicle weight on a stationary tire is static and

FIGURE 7.4 Cord pattern in a radial tire


480 MECHANICS OF PNEUMATIC TIRES

nonsymmetrical, that by centrifugal force on an unloaded spinning tire is


steady state and symmetrical, that by impact on a road obstacle is dy
namic and non-uniform, that by rolling under load is dynamic and cy
clical. In each case the intensity varies with location on the tire. Thermal
loads may arise from unequal expansion or contraction of the rubber and
cord in a tire when the temperature is changed. Both thermal expansion
and Joule effect may be involved even when the temperature change is
uniform throughout the tire. More often the temperature change is non-
uniform, such as that caused by hysteretic heating or sliding of the tread
on a rough surface. Thermally induced forces also arise from cord shrink
age after molding a tire [2]. In addition to their classification as static or
dynamic in nature and mechanical or thermal in source, tire loads may be
axisymmetric or nonaxisymmetric depending on their distribution in the
circumferential direction. Tire load is axisymmetric if its magnitude does
not vary with circumferential location. Examples are loads caused by in
flation, centrifugal forces, and thermal loads from cord shrinkage or uni
form temperature change of a tire. Nonaxisymmetric loads vary with cir
cumferential position, often acting on only a segment of the circumference
such as those in a tire footprint.
Accurate definition of the magnitude and distribution of the loads act
ing on a tire is one of the most difficult aspects of tire analysis. This is re
sponsible for many of the discrepancies between the analytically predicted
and the actual behavior of tires.

7.4 Tire Response


The response of a tire is its reaction to applied loads. This may be a
shape change, a collapse due to instability, or failure.
The change in a tire's geometry is defined by the displacements of se
lected locations on its structure. Strains and stresses at each of these loca
tions are then calculated from their displacements. The remainder of this
chapter describes techniques for determining the displacements, strains,
and stresses in the tire. Comparison of the values of these response func
tions with bounds set on those values will describe the tire's behavior, its
performance, and its durability.

7.5 Analytical Models


Analysis of tires for stresses and deformations can be done at different
levels corresponding to different degrees of detail. One way of distinguish
ing these levels is outlined here and discussed in more detail later in this
chapter. In the following hierarchy each successive level provides a model
that is removed from the exact tire by an additional idealization so that
the solution is more easily obtained.
Cord and rubber model: represents each cord in its exact dimensions and
positions relative to surrounding rubbers. Various rubbers are repre
sented by isotropic, incompressible elements; the bead wires and the
cords are treated as filaments with distinct properties in their longitudi
nal and transverse direction.
Anisotropic ply model: represents separate plies but not individual cords.
It allows for interply strains and nonlinear distribution of strains across
TIRE STRESS AND DEFORMATION 481

the tire thickness. Rubber is treated as in the model above. Beads are
represented by several plies. This model may be analyzed by finite ele
ment techniques with separate elements modeling the different plies
and incorporating their distinct geometries and material properties. Be
cause of the large difference between properties of rubber and the cords,
a preferred representation consists of plies which include the cords and
have thicknesses equal to the cord diameter, and rubber plies in be
tween.
Laminate models: group a number of plies into a laminate with ortho-
tropic properties. Included in these models are:
(a) Finite element models which include elements representing all body
plies, allowing for variation of properties due to variation in cord
spacing depending on radial distance, elements representing the two
belt plies, and elements representing the beads.
(b) Finite element shell elements representing the entire tire thickness
with one element which integrates the constituent plies within the
specific thickness.
(c) Classical anisotropic shells.
Membrane shell: ignores the effect of bending on the tire deformation.
Netting analysis: ignores the contribution of rubber to the load-carrying
function of the tire.
Rings, strings, or beams on elastic foundation models: represent the tire
by equivalent structures.
The simpler models are treated first in order to develop an understand
ing of them and establish the added complexity in moving to the next
model up the hierarchal scale. Also, some of the higher levels of analysis
await further development.

7.5.1 Early Tire Models


Ring and string models and beam-on-elastic foundation models have
been developed and applied extensively by Clark and coworkers [3] and
by other investigators. These models generally involve the use of equiva
lent tire parameters determined from full-scale tire experiments. Analyses
of these models were useful in predicting overall tire characteristics such
as vibration, and responses to side slip, braking, and traction. They pro
vided an understanding of such behavior before the more direct tech
niques were applied to tire analysis. However, the early models were lim
ited because (a) they required extensive experiments for evaluating the
equivalent parameters, (b) their validity was limited to specific ranges of
those parameters, and (c) their domain of validity could not always be pre
dicted in advance.
The early models generally considered the tread to be a prestressed
string (membrane analysis) or a ring (bending included) and the sidewalls
as elastic foundations supporting the tread structure. The more elaborate
models depicted the tire by a sandwich structure whose core resisted shear
deformation while the outer layers resisted bending deformations.
Having acknowledged their early contribution to the understanding of
tire behavior, no further detail is provided here on these models because
of both the aforementioned limitations and the progress made on the more
advanced models. The availability of advanced models has also been in
creasing.
482 MECHANICS OF PNEUMATIC TIRES

7.5.2 Netting Analysis


Analysis of inflated tires is simplified by assuming that inflation pres
sure is supported exclusively by tensile forces in the cords. The supporting
network deforms into an equilibrium shape which depends on the orienta
tion and elastic properties of the cords and on inflation pressure. If the
cords are assumed to be inextensible, the inflated shape becomes inde
pendent of inflation pressure. The analysis involves calculating the stress
resultants in the meridional and circumferential directions, and estab
lishing the tire's profile by relating the radial and axial cylindrical coordi
nates of points on that profile.
A cylindrical coordinate system is used in developing the netting analy
sis. Figure 7.5 shows that the location of a point in space is determined by
its radial coordinate r, axial coordinate z and the angle t) specifying its lo
cation on the circumference of the circular cross-section. Thus a given r
defines a cylinder; r together with z define a circle along that cylinder lo
cated at distance z in the axial direction; and r, z, θ define a point on that
circle. For the tire portion illustrated in Figure 7.6, r represents the dis
tance from the car's axle, z is the distance along the axle, and θ defines the
angular location along the tire's circumference.
Figure 7.7 is a cross section showing a tire profile with radial and axial
co-ordinates r and z. In order to define the profile of an inflated tire by
netting analysis, a relationship must be developed between the coordinates
r and z defining that profile. If the radius of curvature rt is rotated by a
differential of the meridional angle </>, the resulting differential along the
curved profile is rtd<$>. Trigonometry of the differential triangle yields the
following;
dr = rtd<j) cos <#>, (7.1)
and
dz = — sn < (7.2)
In order to relate r to z, these differential equations must be integrated
along the tire's profile. Toaccomplish this, radius of curvature rt must be
expressed in terms of known tire parameters.
Figure 7.6 shows the symbolism and terminology that will be used for

FIGURE 1.5 Cylindrical coordinate system


TIRE STRESS AND DEFORMATION 483

FIGURE 7.6 Cylindrical coordinates in a lire

analyses. Summing the forces that act on the shell segment in the radial
and axial directions yield the following equilibrium equations [4, 5]:

^+^ = />, (7.3)

and

N.- (7.4)
2rsin</>
6
A',,, and Ne are membrane forces per unit length obtained by integrating
their corresponding stresses over the shell's thickness; r^ and r, are the
principal radii of curvature of the shell; p is the inflation pressure; and rw is
the radial coordinate of the widest section on the shell.

FIGURE 7.7 Coordinate differentials on a tire section


484 MECHANICS OF PNEUMATIC TIRES

o M end ion

Mend'cr.

FIGURE 7.8 Cord directions and shell forces in a radial tire

For tires consisting of a radial ply (f) and two bias plies (b) at ±/? de
grees from the meridional direction, the membrane forces may be written

(7.5)
and
(7.6)
Resolving the forces acting on a cord network element bounded by two
meridians and two parallel circles as shown in Fig. 7.8, yields the follow
ing [4, 5]

tan2 y3, (7.7)

= 0, (7.8)
and
(7.9)

where rc is the radial distance from axis of revolution to crown, and α and
/i are tire parameters defined as:

(7.10)

(7.11)
1-y
c - Vi a tan2 & (7.12)
TIRE STRESS AND DEFORMATION 485
Ir

y (7.13)

where ftc is the cord angle at the crown.


^Rguation (7.7) is the basis of netting analysis for composite Structures
consisting of bias plies (b) only. It has been used in designing fabric hoses
[6] and filament-wound rocket motor cases [7] as follows. For pressurized
cylindrical shells the membrane forces are given by [8]:
= prn and

Substituting in eq. (7),


tan2 2;
thus
±54.7° (7.14)
This implies that the cylindrical shape will be maintained upon pressur-
ization only if the cord path is at ±54.7°. For other angles, the matrix ma
terial is subjected to high stresses and the cylindrical shape is distorted by
inflation. For ratios of Ne/Nt different from 2.0, equilibrium cross sections
are not cylindrical. For example, a U.S. patent filed in 1920 [9] recognized
this concept by recommending that tires with large bias angles be cured in
low profile molds and tires with low bias angles be cured in high oval
molds. This process produces tires that are close to their equilibrium con
figuration. The rubber matrix would be relatively free of inflation stresses
and the tires would retain their shape in service.
Using the pantographing law [10, 11]:
r sin ft
(7.15)
rt shift'

and introducing the variable v [4],

(7.16)

the membrane forces may be eliminated from eqs. (7.3-7.9) to give ex


pressions for the equilibrium shape of the tire in the form

(7.17)

and
+ 1v \c + (c + y -
ctn'ft (7.18)
1 + 3? - c + (c + Y - 1) sin <f>
MECHANICS OF PNEUMATIC TIRES

Explicit expression for the variable ν in eq. (7.18) is obtained from the in
verse of eq. (7. 17):
v - $ {1 + 2 [c - (c + y - 1) sin a]} sin2 (i// + 30°) - ±,
where

tt •»" 1 ic - ^ 1- Y- i)smq>\) \^ j

Equation (7. 1 8), with eqs. (7. 1 9) and (7.20), defines the radius of curva
ture % as a function of the meridional angle 4>. Substituting in eqsT(7.1)
Of and (7.2) and integrating the resulting differential equations yields the
radial and axial coordinates at selected values of </>. This defines the tire's
profile. Examples of such tire profiles are given in Figs. 7.9 and 7.10.
For the special case ofbias ply tires, A^0 is zero and therefore, from eq.

(7.21)
Substituting this value in eq. (7. 1 1) yields
c-l-y. (7.22)
Equations (7. 1 7-7.20) then have the following simplified forms:
. (c - v) VI +2*
</> — sin" - '- , (7.23)
c
., _ r,2 c VI +2v
(7.24)

(7.25)

0.3,

FIGURE 7.9 Family of tire profiles with constant crown angle ftc — 58° and rc/rw as a
parameter
TIRE STRESS AND DEFORMATION 487

1.00

OSS
0.05 0.10 O.IS 0.20 025 030
X

FIGURE 7.10 Family of profiles with constant ratio rc/rw •» 0.8 and crown angle f)e as
parameter

and
l+2c\ -3/2
- - \ cos ' c sin<t> (7.26)

The main difficulty in predicting tire shapes arises from the fact that c
and γ are tire parameters whose values are not known before the tire pro
file is known. This is illustrated in eqs. (7.10-7.13) where the values of rr
rc, fic and r^, (c) are needed in order to calculate c and γ. Thus, an analyti
cal solution can be obtained only by a trial-and-error procedure or by as
suming a function for the tire's profile.
Assume an elliptic meridian function where the half-width of the tire is
'a' and the height above rw is 'b', then the tire shape is given by
r-r.
1. (7.27)

Using this function, the radii of curvature and membrane forces in bias
ply tires become

(7.28)

4 + (a* -
T,— (7.29)
a(r ~

(7.30)
488 MECHANICS OF PNEUMATIC TIRES

and

* 2a[b4 + (a2 - 62)(r - /v)2] "2 ' '

Membrane forces of bias-ply tires may also be calculated directly from


eqs. (7.1), (7.2), (7.7), and (7.15), which yield:
prf(r* - rj) cos ft
* 2r(r2-r2sin2ft)1/2' l '

and
v _ /"T,(r2-r2w)cosftsin2ft
Nl 2(r2-r2sin2ft)"2 (733)

The radii of curvature, which may be calculated from their differential


form, are:

and

f> = (r2-^Xr2-Vsin2ft)''2 ' (735)

Netting analysis also can be used to calculate cord tension due to infla
tion pressures [5]:
^r2(r2-r^cosft

where T is the cord force and # is number of cords in the tire. For radial
tires βc = 0, and eq. (7.36) reduces to

r=(r2-r2). (7.37)

Netting analysis can also be used for generating families oftire profiles
[11] for an assumed value of rc and a range of values of ratio rjrc. Such
curves are useful for illustrating the effect of varying a particular tire pa
rameter (e.g. rw) on the shape of the tire's profile.

7.5.3 Limitations of Netting Analysis


Some of the limiting assumptions of netting analysis can be removed
rather easily. For example, cord extensibility can be included in the pant
ograph law of eq. (7.15), thus incorporating the effects of cord properties
and inflation loads on the tire's inflated profile. This has been discussed by
TIRE STRESS AND DEFORMATION 489

Robecchi and Amici [11]. However, the analysis fails to include the stiff
ening effects of the rubber surrounding the cords and of course neglects
the bending of the tire. Both of these are important influences near the
bead area, and in some heavy service tires are important throughout the
structure. Netting analysis has been used with limited success to study
thin-waUecftires, andTepresents a first approximation to the calculation of
tire shape. More detailed knowledge of the tire requires more thorough
analysis.
Although radial plies were included in deriving eqs. (7.5-7.20), netting
analysis cannot be applied to the belt region of a radial tire since the dis
tribution of load between cords in the various plies cannot be determined.
The structure is redundant and eq. (7.7) is not valid in the belt region,
which is the radial tire's most critical area.
Calculating a tire profile by solving the equations of netting analysis
through trial and error can become tedious and demanding, and the solu
tion may not converge. The solution in eqs. (7.28-7.31) for an assumed el
liptic profile may yield values for the membrane force Ne which are differ
ent from those in a non-elliptic tire of the same width and section, due to
small shape differences. The solution also becomes singular at r = rw,
which is the point of maximum width on the tire meridian. This computa
tional difficulty can be removed by following the approach of Nicholson
[96]. This singularity can also be removed completely if a nonlinear mem
brane or a shell bending theory is used in that region.
Finally, while netting analysis can be applied to any known shape, the
calculation of the shape itself has only been carried out for axisymmetric
loads such as inflation or rotation, both of which are of less concern than
deformations in the footprint.

7.5.4 Membrane Analysis


Membrane analysis includes rubber matrix contributions and so is a
step closer to the actual tire than is netting analysis. It does, however, still
neglect the effect of bending on shell deformation and assumes the loads
are carried through stretching a membrane consisting of thin cord-rubber
plies.
The geometric and equilibrium relations for linear membrane analysis
are identical to eqs. (7.1-7.4) developed earlier for netting analysis and, if
an elliptic profile is assumed, radii of curvature and stress resultants of the
membrane are given by eqs. (7.28-7.31). Thus linear membrane analysis
offers no benefit over netting^ analysis since the calculated functions in
both analyses arer independent of the elastic properties of the material,
jhose properties which would reflect the effects of incorporating the rubber
"matrixin the analyzed structure.
However, nonlinear membrane analysis can be used to study non-local
trends for the contact problem. Classical nonlinear membrane theory has
been used in calculating profiles of membranes inflated against flat sur
faces [12, 13]. Nonlinear membrane theory was also the basis of a finite
element approach to deflecting a tire innertube against a flat surface [14].
The latter anlaysis, which was greatly simplified by ignoring both shell
bending and composite properties of the material, was aimed at demon
strating the use of finite elements for more general representation of the
geometry, loading, and boundary conditions in analyzing tire-type struc
tures.
490 MECHANICS OF PNEUMATIC TIRES

7.5.5 Limitations of Membrane Analysis


Nonlinear membrane theory, while providing useful trends, demon
strates the need for incorporating the shell bending stiffness and its com
posite properties in the solution of more general tire problems, especially
that of analyzing tires deflected by footprint loadings.
Membrane analysis is generally limited in scope and application be
cause it:
1. Cannot handle abrupt changes in geometry, material properties, or
loading.
2. Cannot account for the effects of transverse shear deformations in
the contact area.
3. Cannot predict local forces in zones of significant curvature change
or discontinuity in ply orientation.
7.5.6 Thin Shell Analysis
Shell models include the bending effects that are ignored in membrane
theory. Therefore they can account for local effects in the footprint and
bead regions and in regions of significant curvature change.
Classical approaches to shell theory involve solution of differential
equations to obtain closed form expressions which relate the shell's
stresses and deformations to the applied loads as functions of shell dimen
sions and material properties. Broad applications have encouraged contin
ued development of the theories to rather sophisticated levels. However,
most classical shell theories are not directly applicable to tire analysis due
to the tire's complicated shape, variable thickness, nonuniform composite
properties, and large deformations. While many difficulties remain in the
way of a thorough analysis of tire stresses and deformations, thin shell the
ory and finite element modeling are techniques that can be used to obtain
more general solutions for the tire than can membrane theory. These solu
tions involve fewer idealizing assumptions than do those obtained in the
preceding sections.
Thin shell theory was used by Brewer [15] to analyze an aircraft tire
subjected to inflation loading. The tire was modeled as a thin laminate
consisting of cord-rubber plies stacked in a specified sequence. The basis
of Brewer's analysis and some of his results are shown here to present an
approach that is intermediate between the simple but ingenious netting
analysis and the finite element techniques discussed in later parts of this
chapter. .
Thin shell theory assumes:
1 . Shell thickness is small relative to the radii of curvature of the shell
surface.
2. Displacements are small relative to shell dimensions. (Effect of large
displacements are accounted for by applying the loads incrementally
and updating the shell geometry.)
3. Stresses normal to the surface are negligible.
4. Sections through a shell's thickness that are planes before deforma
tion of the shell remain planes after deformation.
Equations of equilibrium for a differential shell element are derived, as
in netting analysis, by adding the forces and moments along the element's
axes [15]. These equations lead to the following six first order differential
equations in fundamental variables u, w, «4, Q+, N+, and Mt for an
axisymmetric shell.
TIRE STRESS AND DEFORMATION 441

1 du N+ I A I2cos«j> a sin<>
u— + — (7.38)
7.~dt"^~\ Anr
1 du>4, _ 1 )12 costjt
(7.39)

1 dQ. co$4>
An r
sin<f>costj>

I An Aj2 ~ Atl
(7.40)
1

J^ dM. _ costj> IDtl- Dtl

(7.41)

1 dw 1
(7.42)

and
cos<>
r* d$ An
N.--Q.
COStj)

(7.43)
\ *

where </„, </,.. are the external loads on the shell, u is the tangential dis
placement, w is the normal displacement, u>» is the rotation, Q.., is the trans
verse shear force, N+ is the membrane force, M+ is the bending moment,
and Av and D,, are tire laminate stiffness properties calculated from:

(7.44)

and
i ",
Is (7.45)
492 MECHANICS OF PNEUMATIC TIRES

The (Qij)k are elements of stress-strain matrix Q of the Arth ply in the lami
nate, located at distance hk from the laminate's midplane.
The general relationship between a composite laminate's forces and mo
ments and its curvatures and midplane strains is given by:

N.
BK (7.46)
Bn BU BI6 Dn

M, B2I B22 B26 D2i K.


BI, B62 BM Dbt

where e£, e0,, yj are strains at the shell's middle surface, and A*, K» K^ are
changes in curvature and torsion. The constitutive relation for stretching is
provided by matrix A, that for bending by matrix D, and that for bending-
stretching coupling by matrix B.
Equations (7.38-7.43) are based on a number of simplifying assump
tions. If the laminate is assumed to consist of plies with cords oriented at
+θ and -θ angles, stacked symmetrically relative to the laminate's mid-
plane, then several elements of matrices A, B, and D vanish.

^16 Atl — AM — /462 — 0


Dtt = £>„, = D26 = D62 = 0 (7.47)
^n = ^12 = B2I = B22 = B,*, = 0
If it is further assumed that the laminate is made up of a large number
of plies (more than four plies),
B16 = Btl = B26 = B62 = 0, (7.48)
and eq. (7.46) is reduced to:

". Alt AI2 0 <


N. = A2t A-a 0 4 (7.49)
N« 0 0 A* ri
M. Du D12 0 K.
M, = DH D22 0 K.
0
1 * --r1 0 />„ K«
' Only the above remaining nonzero elements of matrices A and D appear
in the differential equations (7.38-7.43).
TIRE STRESS AND DEFORMATION 493

Equations (7.38-7.43) and the prescribed displacement values at the tire


boundary define a boundary value problem. Following techniques de
scribed by Kalnins [16], the problem is replaced with a set of initial value
problems that are solved by the Runge-Kutta numerical integration ap
proach. To insure numerical accuracy in solving the resulting simultane
ous equations [16], the meridional length of a shell with thickness t and
radius of curvature r must not exceed the value 3/λ, where
3(1 - v* v*)

This criterion is met by dividing the shell into a number of segments and
performing stepwise numerical integration on each segment. Shell seg
ments may be assigned differing thicknesses and material properties, de
pending on their locations on the tire's meridian. The load is applied in in
crements that are chosen to produce "small" displacements.
This procedure was used to analyze a six-ply aircraft tire for a 655 kPa
inflation pressure. Shell displacements, rotations, membrane forces, mo
ments, strains, and cord loads were calculated at 22 stations on the tire's
half meridian. The results are illustrated in Figures 7.1 1-7.20. Good corre
lation was obtained between calculated and experimental values for in
flated profiles, Figure 7.1 1, and for strains in the crown area, Figure 7.18.
Profiles of the calculated tensile membrane forces, Figure 7.12 and bend
ing moments, Figures 7.14 and 7.15 are as expected. However, the calcu
lated nonzero values for membrane shear force N+e, Figure 7.13 and twist
ing moment A/,,,/,, Figure 7.16 violate assumptions used in the analysis,
although these results may be due to computational difficulties rather than
the formulation. The significant calculated value of shear force Qv Figure

655 kPa (CALC.)


140 EXPERIMENTAL
0 kPa

20 40 60 80 100
AXIAL DISTANCE, mm
FIGURE 7. 1 1 Meridian profile of inflated 32 X 8.8, 6-ply aircraft tire at 655 kPa inflation as
determined experimentally and as calculated by thin shell analysis
494 MECHANICS OF PNEUMATIC TIRES

2.4

FIGURE 7.12 Calculated tire membrane tensile forces as a Junction of position on the
meridian of an aircraft tire at 655 kPa inflation pressure

7.17, and the slow decay of moments M+ and Me, Figures 7.14 and 7.15,
show a clamped edge effect at the beads and a violation of the true thin
shell phenomenon.

7.5.7 Limitations of Thin Shell Analysis


Limitations of the thin shell approach described above include:
1 . KirchhofFs hypothesis of plane sections remaining plane after defor
mation is not strictly valid for tires because of the differing deforma
tions of different plies.

FIGURE 7.13 Calculated membrane shear force as a function of position on the meridian of
an aircraft tire at 655 kPa inflation pressure
TIRE STRESS AND DEFORMATION 495

BEAD
2.4

£-50
DO

FIGURE 7. 14 Calculated lateral bending moment as a Junction ofposition on the meridian of


an aircraft tire at 655 kPa inflation pressure

2. Assumptions used in reducing constitutive eq. (7.46) to the simplified


form of eq. (7.49) has serious implications: (a) tire plies are normally
stacked with alternating cord angles, resulting in laminates that are
not symmetric about their midplanes, and (b) passenger car tires nor
mally include no more than four plies. Therefore, the B matrix does
not vanish and its bending-stretching coupling effects can be quite
significant.
3. The simple support condition used in the analysis neglects the forces
at the tire-rim interfaces that are needed to prevent tire warping.
4. Calculated profiles for the example aircraft tire agree well with ex
perimentally obtained inflated profiles. However, some strains, shear
forces, and moments correlate poorly with measurement and violate
assumptions used in deriving the theory.

FIGURE 7.15 Calculated circumferential bending moment as a function of position on the


meridian of an aircraft tire at 655 kPa inflation pressure
496 MECHANICS OF PNEUMATIC TIRES

_ 50

? 30

,0
CROWN

o -10

1-30
co

FIGURE 7.16 Calculated twisting moment as a function of position on the meridian of an


aircraft tire at 655 kPa inflation pressure

5. The incremental finite difference solution is time consuming; one


must resort to trial and error to determine the admissible size of each
load increment.
6. The approach is limited to analyzing the inflation or rotation prob
lems. Analysis of tires deformed by other loadings remains beyond
its scope.
7.5.8 Finite Element Models
In the finite element approach to structural analysis, structures are mod
eled by a number of interconnected substructures called elements. The
idealization is such that the response of each model element is known

-10+

FIGURE 7. 17 Calculated transverse shearforce as afunction ofposition on the meridian ofa


aircraft tire at 655 kPa inflation pressure
TIRE STRESS AND DEFORMATION 497

\ EXPERIMENTAL

(LATERAL)

FIGURE 7.18 Calculated (solid line) and experimental (points) tire membrane strains as
Junctions ofposition on the meridian of an aircraft tire at 655 kPa inflation pressure

(stress or deformation due to load), and the response of the entire structure
is determined by "adding up" the response of its elements. In the limit, as
the "finite" size of each element is reduced to zero, the model and the orig
inal structure become identical.
Although the finite element method did not come into use until the

.08
(LATERAL)

-.02

FIGURE 7.19 Surface strains as functions ofpositions on tire meridian of an aircraft tire at
655 kPa inflation pressure
498 MECHANICS OF PNEUMATIC TIRES

28 OUTER PLY

MIDPLY

1.2 1.6 2.0 2.4

FIGURE 7.20 Collated cord loads asfunctions of location on the meridian ofan aircraft tire
at 655 kPa inflation pressure

1950's [29], its basis of representing a continuum by a set of discrete ele


ments has been in use much longer. Most portal and gable frames ana
lyzed as assemblies of beam and column "members" are actually continu
ous at their joints. In 1941 Hrennikoff [17] introduced a theory for
representing arbitrarily shaped plates by beams along their edges and di
agonals. Variable-thickness plates were modeled as planar frames, and
curved shells as space frames. These frames were then analyzed by avail
able analytical techniques.
When constitutive equations are written in matrix form they can be
solved efficiently on a computer by matrix algebra methods [18]. One of
the following methods may be used to formulate the problem.
1. Displacement method: Displacements are determined by solving
"equilibrium" equations. The coefficients in those equations consti
tute the "stiffness' matrix.
2. Force method: Forces are determined by solving "compatibility"
equations. The coefficients in those equations constitute a "flexibil
ity" matrix.
3. Mixed or hybrid method: Both displacements and forces are deter
mined. The coefficient matrix includes both stiffnesses and flexibili
ties.
The approach best suited for a given problem depends on the structure's
geometry, loading, and support conditions. These factors influence the size
of the problem and the ease of its solution in each approach. The dis
placement method, the one most widely used with finite elements, is illus
TIRE STRESS AND DEFORMATION 499

trated by an axially loaded spring. Its load-displacement relation is given


by:

where P is the applied load, δ is the resulting displacement, and k is the


spring stiffness, i.e., the load required to cause a unit displacement.
Numerous finite element models have been derived, offering a variety of
formulations and applicabilities. The development and general character
istics of two elements representing extreme levels of sophistication are
shown below in order to illustrate their capabilities and applications and
to point out bases common to them and to almost all element models.

7.5.8.1 Constant Strain Triangular Element


The basic features of a constant strain triangular element are its simple
shape, which enables it to represent complicated geometries without sig
nificant deviation from the exact geometry, and its linear displacement
functions which simplify derivation of the element properties. Figure 7.21
illustrates a segment of the toroidal shell element defined by the corner
nodes i, _/, and k of its triangular cross section. If it is assumed that the
loads are axially symmetric and that the material is either isotropic or or-
thotropic, the displacements of a point in the r-z plane may be defined [19]

u, - b{ + b-i r +
(7.52)
u, = bt + b, r +
where ur is the radial displacement, uz is the axial displacement bi are the
(undetermined) coefficients, r is the radial coordinate of the point, and z is
the axial coordinate. This linear displacement variation within the element
corresponds to constant element strains since the latter are first derivatives
of the displacements.

FIGURE 7.21 Toroidal finite element having triangular cross-section. Nodal positions art
determined by the two coordinates r and z.
500 MECHANICS OF PNEUMATIC TIRES

7.5.8.2 Element Stiffness Matrix


From eq. (7.52) the displacements of the corner nodes are given by:

000

1 rt z, 0 0 0

«,* 1 rk zk 0 0 0
u.1 (7.53)
0 0 0 1 rt z, ft.

00 0 1 rt Zj

00 0 1 rk zk

where a bold faced symbol designates a matrix notation and a superscript


on a displacement component designates the referenced node. Inverting
the square matrix on the right hand side of eq. (7.53) yields:
b = hu« (7.54)
which relates the undetermined coefficients b to node displacements ue.
Assuming the strains to be small, they can be calculated from:

€ -
" dr

(7.55)

€" dz + dr

Performing these differentiations on the displacements of eq. (7.52) yields


the following strains:
g b, (7.56)
where

^ 0 1 0 0 0 0
«M 0 0 0 0 0 1
, and g = (7.57)
CM 1 z/r
l/r 0 0 0
€„ 0 0 \ 0 1 0
TIRE STRESS AND DEFORMATION SOI

Combining eqs. (7.54) and (7.56) yields


u. (7.58)
or
B-g-h, (7.59)
where B is the strain-displacement matrix.
Strain-energy principles yield the following form for the element stiff
ness matrix:
k - / B' • C • B dv, (7.60)

where k is the element stiffness matrix, C is the stress-strain matrix, ν is the


element volume, and a t superscript on a matrix designates the transpose
[18] of that matrix. Substituing eq. (7.59) into eq. (7.60) gives
k = h' s h, (7.61)
where
- I g1 • C • g dv. (7.62)
»»

For orthotropic materials such as those formed by radial plies or by


pairs of bias plies, the stress-strain matrix C can be written as:

CM C» C13 0
c« C
^•22 C
^23 0
(7.63)
c» C23 C33 0
0 0 0 c«
and the product matrix inside the integral sign in eq. (7.62) is
g' C g
1 i«.*« 0 0
1
P-
0 0 (C12 -
z
" ) 0 C

0 0 0
Symmetrical
0

(7.64)
502 MECHANICS OF PNEUMATIC TIRES

Thus the matrix s can be calculated by integrating each term in the above
matrix over the element's volume.
Closed form expressions may be derived easily for each term in s as a
function of matrix C and the coordinates of the corner nodes. However,
since the calculations involve small differences between large quantities
(due to the small size of the triangle relative to the coordinates of its
nodes), the results become extremely sensitive to computer roundoff er
rors. To avoid this problem, numerical integration formulas have been de
rived [20, 21] which convert the integrals to calculations of areas within
the triangles and various moments of those areas around the reference
axes.
In numerical integration, the definite integral of a function f(x) is ap
proximated by a properly weighted sum of particular values of f(aj) with
the abscissas aj suitably distributed within the limits of integration. The
abscissas may be preassigned or unassigned. The name Gaussian quadra
ture is associated with situations where the sample points aj are unas-
signed; the sample points are all distinct, real, and lie in the interval of in
tegration. The weights are all positive. Certain advantages of numerical
integration and the resulting simplification of the mathematical ex
pressions have been pointed out by many sources including ref. [39].

7.5.8.3 Solution
Taking advantage of the symmetry of the tire's cross section about the
radial axis r, only half of the section is modeled by a finite-element grid.
The grid illustrated in Figure 7.22 includes nodes at material interfaces so
that each element is entirely within one of the tire's constituents, and is
thus represented by the stress-strain matrix of that constituent.
The stiffness matrix for the total structure is obtained by adding the
stiffness matrices of the elements. The matrix rows and columns corre

FlGURE 7.22 Finite element grid for bias ply tire


TIRE STRESS AND DEFORMATION 503

spending to each node derive their values from the stiffness matrices of the
elements connected to that node. Thus,

K= 2k, (7.65)

where K is the structural stiffness matrix, kl is the stiffness matrix of the lth
element, and «.. is the number of elements in the finite element grid. Struc
tural equilibrium is defined as in eq. (7.51) by
K8 = P (7.66)
where δ is the matrix consisting of the displacements of the nodes in the
finite element grid, and P is the matrix of the loads applied at the nodes.
Inverting eq. (7.66) gives
8 = K-'P . (7.67)
which defines the displacements of the nodes. The strains in each element
are then calculated from eq. (7.58) by selecting the components of δ which
form ue of that element, and using matrices g and h which were formed in
deriving the stiffness matrix.

7.5.8.4 Applications
Constant strain elements of triangular cross section have been used [2]
to analyze the deformations of bias tires caused by cord shrink forces.
These forces develop during tire cure when temperatures are high enough
to allow recovery of some of the frozen-in stresses introduced during
drawing of the filaments or "treatment" of the fabric. The result may be
distortion of the tire when it is removed from the mold.
Given the composite properties of the green tire, the pantographing law
and the shaping equations [11] define the orientation and spacing of the
cords in each element depending on its location on the tire's profile.
Shrink forces along the cords are transformed to cylindrical coordinates r,
z, θ to determine a set of "initial" stresses a", a", <ft and T". for each element.
Using energy principles, an equivalent load matrix is derived for each ele
ment such that the work done by the equivalent nodal forces acting on the
corresponding node displacements is equal to the work done by the initial
stresses acting on the corresponding strains. The following equivalent
loads are derived:
Or' (z, - zk) (rk - r,) of
Qi = it (zk - z,) (r, - rk) (7.68)
Q,k
and
(zj-zk) (rk-rj)
01 irr (zk - z,) (r, - rk)
Q." (z, - Zj) (r, - r,)
504 MECHANICS OF PNEUMATIC TIRES
MOLDED
FINAL

360

180 mm

FIGURE 7.23 Deformation due to shrink forces; initial molded shape and computer-drawn
calculatedfinal shape

where r and z are the coordinates of the centroid of the triangular element.
Combining the equivalent loads of all the elements generates the load
matrix P which is used in eq. (7.67) to calculate the nodal displacements δ.
These displacements then define the deformed tire profile and provide a
comparison between the molded shape and the shape of a demolded tire,
as illustrated in Figure 7.23. Good correlation was obtained between cal
culated and experimental results.

7.5.8.5 Nonaxisymmetric Analysis


If loads are applied non-uniformly at different points on the circum
ference of the tire (e.g., footprint loads) or if the material is anisotropic
(e.g., a single belt ply), then the tire's deformation is not axisymmetric. Its
response is influenced in three ways.
1. Besides their axial and radial displacements, points on the tire's cross
section undergo a third displacement—in the circumferential direc
tion. Thus, the following component is added to equation (7.52):
u, = b7 + bsr + (7.69)
2. Two additional strains erB and e,.» are added to the strain matrix.
These strains, which were suppressed in the axisymmetric problem,
represent torsional deformations; their corresponding stresses are il
lustrated in Figure 7.24.
3. The displacements and strains vary around the circumference and
become functions of the angular coordinate.
The axisymmetric element previously formulated can be modified for
analysis of certain nonaxisymmetric problems without resorting to a full
three-dimensional analysis. If the load variation in the circumferential di
rection is symmetric about a plane containing the axis of revolution r,
these loads can be expanded in Fourier series as
R = £ &»(r, 2) cos n6,
Z = £ Z,(r, z) cos n6, (7.70)
TIRE STRESS AND DEFORMATION 505

FlGtJRE 7.24 Shear stresses in nonaxisymmetric problems

and 9 r, z) sin n9,


where R, Z and ft are radial, axial, and wind-up components of load ma
trix P, and the subscript n refers to the amplitude values for the nth har
monic (i.e., nth term in the Fourier series expansion).
The Fourier expansion of a function f(θ) may be written as:
f(0) = aa + a, cos 0 + a2 cos 20 + 1- a, cos nff, (7.71)
and the coefficients are calculated from:

and

a, = — cos nOd6.

By making use of the orthogonality of harmonic functions, the three-di


mensional problem is thus divided into a series of uncoupled two-dimen
sional analyses.
In each analysis, formation of the stiffness matrix follows eqs. (7.52-
7.62) and the resulting displacements and strains are calculated as in the
basic triangular element. The results are then combined as follows to de
termine the displacements and strains and their variations in the circum
ferential direction:

«, = Z «,„ cos nO
u, - £ u!n cos nd
"» = £ "»„ sin nO
e^-S^, cos n8 (7.73)
e,, = S Cx» cos n6
€«. = !€„,,, cos n8
S €,„ cos nO
2 e,*, sin nB
and S «zfc sin n6.
506 MECHANICS OF PNEUMATIC TIRES

7.5.8.6 Applications
Fourier decomposition was used in ref. [22] to analyze stresses and de
formations of a deflected tire using the finite element grid shown in Figure
7.25. The inflated profile and the deflected profile were calculated. The re
sults plotted in Figure 7.26 show satisfactory agreement between the mea
sured profile and the calculated profile of the deflected tire. The analysis
also yielded values for cord tensions in the center of the footprint and at
90° from that location. These values, as illustrated in Figure 7.27, show
good agreement with experimental results. The calculated distribution of
the principal shear strain in the outer belt is illustrated in Figure 7.28. The
broad range of results obtained from this anlaysis demonstrates the
strength of the finite element approach in predicting tire stresses and de
formations. Higher order elements should provide even better details and
more accurate results.

7.5.8.7 Limitations of Constant Strain Models


The constant strain element of triangular section contributed signifi
cantly to the understanding and application of finite element techniques.
It continues to offer a simple and flexible tool for analysis of a broad class
of problems. However, its usefulness in analysis of tire stresses and defor
mations is limited by the following.
1. The constant strain feature requires a greater mesh refinement in
modeling the tire than is needed with higher order elements. This has
two undesirable effects:
a. More nodes means more variables to determine. A larger coeffi
cient matrix leads to greater computation costs and a reduced nu
merical accuracy for the solution.
b.While the number of elements does not influence the size of the co
efficient matrix, its undesirable effect on numerical accuracy hat
been demonstrated [23].

FIGURE 7.25 Finite element grid for analysis of tire deflection


TIRE STRESS AND DEFORMATION SOT

UNLOADED PROFILE
UNINFLATED
k — INFLATED TO 216 kPa
\ (CALCULATED)
I I LOd
M --CALCULATED
'f. MEASURED

FIGURE 7.26 Comparison of calculated and measured tire shapes on the loaded tire. The tire
was deflected 17.3 mm at its footprint center by loading it against a 1707 mm drum

2. The shape and orientation of the triangle can also influence the re
sults. To improve numerical accuracy the length ratio of the sides
must be limited and the triangles must be oriented in a preferred di
rection [24].
3. The triangular element derived above is limited to linear theory.
Even if the tire profile is updated with each load increment, a signifi-

-200
-CALCULATED
— EXPERIMENTAL
FIGURE 7.27 Comparison of calculated and measured cord forces at the crown in the outer
bell ply of a I75SRI4 stationary tire inflated to 216 kfa and loaded with 2.94 kN
508 MECHANICS OF PNEUMATIC TIRES

— «— 60O mm DRUM 01 A.

-o— 1707 mm DRUM DIA.


0.30
-*- FLAT PLATE
2 0.20
l-
co

10 20 30 40 50 BELT EDGE
MERIDIONAL DISTANCE ALONG BELT
(mm)
FIGURE 7.28 Calculated principal shear strain along outer belt ply of a 175SR14 tire inflated
to 216 kPa and loaded with 2.94 kN

cant nonlinearity due to inflation pressure is neglected. Inflation


pressure stiffens the tire and adds so-called "geometric stiffness" to
the linear stiffness of the elements.
4. In deriving the element's stiffness matrix, the formulation of stress-
strain matrix C in eq. (7.63) in effect neglects the bending-stretching
factor reflected in eq. (7.46). This underestimates certain tire stiff
nesses, which may lead to unrealistically high displacements.
5. The application of constant-strain triangles to analysis of tire defor
mations for axisymmetric loadings due to inflation pressures and
cure shrink forces is straightforward. However, analysis by Fourier
decomposition becomes cumbersome. When loads are distributed
over a small segment of the circumference, a large number of terms
must be included in the Fourier series which increases computation
costs and the times needed to prepare the input information and to
interpret the results.

7.5.8.8 SheU Elements


A thin shell element extending through the tire thickness was used by
Zorowski and Dunn [25] for analysis of bias ply tires subjected to inflation
pressures and centrifugal forces. A typical finite element grid used in the
analysis is shown in Figure 7.29. The element's composite properties were
modeled similarly to those in the procedure used by Brewer [15]. No solu
tion is given for the deformation of tires due to inflation loading only. Al
though a fair agreement was obtained with the experimental results of
ref. [26] for a rotating inflated tire, the solution technique had a number
of serious drawbacks. Convergence of the iterative process and accuracy
of the solution were dependent on the size of the element and on the mag
nitude of loading increment. Thus the anlaysis does not insure in advance
that a given finite element grid will lead to an accurate solution, or that a
selected loading sequence will converge. The element model was im
proved in ref. [27] by modifying the iterative solution technique and re
calculating the tire's stiffness matrix at selected load increments to incor
porate the geometric stiffness. Figures 7.30 and 7.31 respectively show the
agreements obtained between calculated and measured profiles for a tire
subjected to inflation pressure alone and coupled with centrifugal forces.
TIRE STRESS AND DEFORMATION 509

NODE i
ELEMENT i

FIGURE 7.29 Gridfor shell elements

The analysis was extended, by use of Fourier expansion, to analyzing


bias ply tires subjected to footprint loads. Dimensions of the footprint and
the distribution of contact forces within the footprint were calculated by
trial-and-error and then used as input to the analytical program. Figure
7.32 shows the deflected profile in the center of the footprint for a tire
without a tread. The results are in good agreement with those measured in
ref. [28]. Figures 7.33 and 7.34 show variation of the deflection and cord
force in the circumferential direction. These results reflect a concentrated
deformation at the footprint (less than 40° sector of the circumference)
and a steep change in cord force magnitude within and near the footprint
zone. Figure 7.35 indicates significant changes in tire cord forces due to
tread removal.

7.5.8.9 Limitations of Shell Elements


The positive features of the modified curved shell element include (a) its
representation of the tire composite properties incorporating bending-
stretching coupling effects, and (b) its accounting for geometric non-

100
E
E80

£60f
FINITE ELEMENT
*40f — (REF. 27)
• • EXPERIMENT
(REF. 26)
UJ

H 200 240 280 320


RADIAL DIMENSION, mm
FIGURE 7.30 Comparison of tire profile calculated by finite element method with that
determined experimentally. Inflation loading only, 165 kPa inflation pressure in 8.25-14 2-ply
polyester bias tire
510 MECHANICS OF PNEUMATIC TIRES

FINITE ELEMENTS
£40- (REF. 25)
• MEASURED
FINITE ELEMENTS
(REF. 26)
20 \
(REF. 27)
X BEAD CROWN
< •l—I
220 260 3OO 340
RADIAL DISTANCE, mm

FIGURE 7.31 Comparison of tire profiles calculated by finite element method with that
determined experimentally. Inflation loading at 165 kPa plus centrifugalforces
produced at 1600 rpm 8.25-14, 2-ply polyester bias tire

linearities in the element-stiffness matrix. However, the shell element re


mains limited by the following:
1. Special care is needed in preparing a finite element grid. Location of
the nodes on a smooth profile and within specific arc lengths is nec
essary for accurate computations. Convergence cannot be taken for
granted.
2. Although inflated profiles are accurately determined, calculated
stress magnitudes may be questionable, especially in areas of high
stress gradient and at ply discontinuities.
3. Extending the technique by Fourier expansion becomes rather cum
bersome; the analysis already includes a large number of loading
steps and several equilibrium iterations within each step.
4. The scope of application is also restricted by the element's limitation
to orthotropic material properties.

100
E
E 80

60 CALCULATED
• «(REF. 27)
40 MEASURED
(REF. 28)

20

180 200 220 240 260 280 300 320


RADIAL DIMENSION, mm
FIGURE 7.32 Comparison of the profile at center offootprint for a aetreaded tire calculated
byfinite element with that determined experimentally. Inflation loading at 207 kPa plus 3.1 kN
radial load
TIRE STRESS AND DEFORMATION 511

0 (DEGREES)

DEFLECTEDV
PERIPHERY

-20-.o o Jo 20
FIGURE 7.33 Calculated circumferential profile of crown of tire under 3.1 kN load and 207
kPa inflation pressure

5. Assuming that sections through a tire thickness that are planes before
deformation of the tire remain planes after deformation, coupled
with representing the entire tire thickness by a single element, ig
nores differential ply responses which may result in significant inter-
ply stresses.
7.5.8.10 Isoparametric Elements
Constant strain elements of both triangular and quadrilaterial shapes
have been modified and refined extensively since they were first developed
in order to improve their capability of modeling complicated configura-

28-

-40 -20 0 20 40
CIRCUMFERENTIAL ANGLE,
DEGREES
FIGURE 7.34 Calculated tire cord forces, inflation loading at 207 kPa plus a 3.1 kN radial
load
512 MECHANICS OF PNEUMATIC TIRES

-40 -20 0 20 40
CIRCUMFERENTIAL ANGLE,
DEGREES
FIGURE 7.35 Calculated cordforces in detreaded lire. Inflation loading at 207 k Pa plus a 3.1
kN radial load

tions. Refinements included using higher order displacement functions to


allow variation of strains within the element, incorporating nodes in addi
tion to the element vertices, and extension to three-dimensional (i.e., solid)
elements.
Early finite element models for analysis of three-dimensional structures
were tetrahedrons [30] and rectangular prisms [31] which constituted ex
tensions of the basic two-dimensional models. In spite of numerous refine
ments these elements remained difficult to implement, and required a
great deal of preparation and computational efforts. The most significant
breakthrough [32] came with the introduction of isoparametric hexahedral
elements [33, 34]. Generalized hexahedral elements are obtained through
the transformation of a hypothetical unit cube, shown in Figure 7.36, to an
arbitrarily distorted six-sided solid shown in Figure 7.37. The dis
placement function is defined in the "natural" coordinate system ξ, T/, ζ in
which each coordinate axis is associated with a pair of opposing faces hav
ing coordinate values ±1. These coordinates correspond to non-dimen
sional cartesian coordinates in the "parent" hypothetical cubic element.
In isoparametric elements the geometry of the general hexahedron is de
fined by the same function that defines displacements within the element.
Subparametric elements [35, 36] offer further advantages for tire-type
structures by defining the geometry with a lower order function; the geom
etry of any tire segment can be adequately represented by quadratic poly
nomials. Derivation of the subparametric element developed in ref. [35] is
presented here in order to show the basis of this family of elements, whose
application to analysis of tire stresses and deformations is expected to in
crease in the future due to their strong advantages.
TIRE STRESS AND DEFORMATION 513

FIGURE 7.36 Parent cube for hexahedral element

7.5.8.11 Element Description


The curvilinear form of the element will be of the second order while
the displacement function will be a fifth order polynomial. To define
quadratic lines, three points are needed. Thus one midside node is re
quired per edge. The eight corner nodes and twelve midside nodes are
shown on the parent cube of Figure 7.36. When the cube is distorted to a
curvilinear form, as illustrated in Figure 7.37, the nodes maintain the same
local coordinates as in the parent cube. Their coordinates in the cartesian
space x y z define the geometry of the element as it is in the structure.
The local coordinates of the nodes as given in Table 7.1 are maintained
for all the elements. This reduces the required input and makes possible
the use of certain topological matrices throughout the calculations once
they are developed at the beginning of the solution. The entire geometry
of a given structure is defined by specifying the "global" cartesian coordi
nates, x, y, z of all the nodes in the finite element mesh.
Midside nodes are introduced only for the definition of the geometry.
The unknown displacements correspond to the corner nodes. There are
twelve unknowns per node, the three translations u, ν, w in the x, y, z di
rections and three slopes per displacement. The definition of slopes at the
corner nodes is better than defining the displacements at the midside
nodes due to the higher valency of the corner nodes. Valency, a name bor
rowed from chemistry, is the number of elements connected to a particular
node.
7.5.8.12 Displacement Field
In the absence of body forces, the displacement u must satisfy the fol
lowing equilibrium equation in the parent cube [38]. Displacements ν and

FIGURE 7.37 Distorted hexahedron


514 MECHANICS OF PNEUMATIC TIRES

TABLE 7.1- Curvilinear coordinates ofthe nodes ofthe isoparametric element in Figure (7.37)

Local Coordinates
Node £
1
2 -
3
4
5 -
6 - -
7 —
8
9 <1
10 ()
II (> -
12 ()
13 (}
14 <»
15 <>
16 <>
17 - <)
18 - <>
19 (>
20 1 <)

it1 have the same formulation as the one given here for displacement u:
V4u = 0, (7.74)
where
a4 a4 (7.75)

For the above biharmonic to vanish throughout the elastic body, the ex
pression for u should not include terms higher than the third power in any
variable (e.g., £*, £*) or terms higher than the first power in any pair of
variables (e.g., fTj2, ξ37/2). Elimination of such terms from a complete fifth
order polynomial results in the following displacement function:
/ = 1 1, fc n, r, ?. a , r, e, es,
r, ?rj, (7.76)

A similar function is used in ref. [36]. There are 32 terms in [f] which cor
respond to 32 .-/-displacement variables: the (/-displacement and its three
derivatives for each of the eight corner nodes. The u-displacement may
now be written as
u - in bu) (7.77)
where h is the matrix of displacement constants. Similar matrices b , and
TIRE STRESS AND DEFORMATION 515

bx define the ν and w displacements, and the total coefficient matrix is

(7.78)

Equation (7.77) is analogous to eq. (7.52) which defined the displacements


for the constant strain triangular element. For further derivation of iso
parametric elements the reader is referred to refs. [33], [34], and [98].

7.5.8.13 Two-Dimensional Analysis


Isoparametric elements also offer a powerful and efficient tool for analy
sis of two-dimensional and axisymmetric structures [33]. Due to their high
order displacement functions and curved edges, the needed degree of grid
refinement is greatly reduced. This leads to a better numerical accuracy in
the solution [37], and reduced computation costs and pre- and post-proc
essing efforts.

7.5.8.14 Variable Material Properties


One of the advantages of using the above formulation for the isopara
metric family of elements is that variation in material properties within
the element can be allowed without unduly complicating the analysis [40].
The variation in properties may be due to changes in material (e.g. differ
ent plies and different rubbers in the tire) or due to changes in moduli
caused by different stress levels (e.g. different degrees of inelasticity).
For a two-dimensional element, the three-point Gaussian quadratures
are shown in Figure 7.38. These correspond to 27 integration points in the
three-dimensional version of that element. Different material properties
may be used at each indicated point depending on its location in a given
tire ply or rubber material.

FIGURE 7.38 Three-point gaussian quadrature points


516 MECHANICS OF PNEUMATIC TIRES

°i>o us fio tie sta ess no ten 300 JIB 390 3« MO

FIGURE 7.39 Isoparametric finite element grid for HR78-15 tire

7.5.8.15 Applications
Results of recent applications of two-dimensional isoparametric ele
ments to tire analysis are described in the following paragraphs. The first
application of three-dimensional elements to calculation of stresses and
deformations due to footprint loading will be awaited with interest.
7.5.8.16 Tire Inflation
Axisymmetric isoparametric elements were used for analysis of an
HR78- 1 5 tire for stresses and deformations due to inflation pressure. The
finite element grid shown in Figure 7.39 consists of two groups of ele
ments: a linear group of isotropic elements representing the rubber stocks
in the tire, and a nonlinear group of orthotropic elements which represent
the belt plies, body plies, turn-up and insert plies, and the bead.
The elastic constants used in the analysis are given in Tables 7.2 and
7.3. Growth in the radial tire's cross section due to 165 kPa inflation is
shown in Figure 7.40. Displacements were outward at the tread and lower
sidewalk and inward but smaller at the shoulders and upper sidewalls.
The calculated outward displacement of the crown was 0.32 mm.

TABLE 7.2 Input propertiesfor linear elements of HR78-15 tire


Modulus of Poisson's
Elasticity Ratio
E (MPa) r
Tread Rubber 4.14 0.49
Bead Filler 6.89 0.49
Other Rubber 6.21 0.49
TIRE STRESS AND DEFORMATION 517

S.
-"" £

&• ON »

odd

u
— r >: I

~ 3 J a: «
fPll'
ma HOT i
518 MECHANICS OF PNEUMATIC TIRES

~-'i»o-']2o-iao-'»o -to -40 -20 o' 20 40 BO «S ibo lio i'«o


2 (MM)

FIGURE 7.40 Calculated HR78-I5 tire section growth due to 165 kPa inflation pressure

Four-point Gaussian quadratures were used in numerical integrations


of nonlinear elements; thus sixteen sets of stresses were calculated for each
element. Stresses in the belt elements were resolved along the cords (at
±22° from circumferential axis) and converted to cord forces. The calcu
lated cord forces due to 165 kPa inflation were plotted in Figure 7.41; they
show a decline of cord load magnitude in the outer belt from a 29.8 N
value at the crown to 2.4 N at the edge. Cord forces in the first belt are
somewhat higher than those in the outer belt, and drop to almost zero in
the step.
Axisymmetric analysis may also be used to determine the effect of a
tire's aspect ratio on the direction of its displacements due to inflation. For
a simplified isotropic material, analysis yields the results shown in Figures
7.42 7.44 which demonstrate that toroids deform towards a circular cross
section upon inflation.

10 20 30 40 50 60 70 80
DISTANCE ALONG BELTS, mm

FIGURE 7.41 Calculated variation of cordforces in belt plies of an HR78-15 tire at 165 kPa
inflation pressure
TIRE STRESS AND DEFORMATION 519

40 -Vo a i5~
H (MM)

FIGURE 7.42 Calculated deformation pattern due to inflation 165 kPa pressure. Tire aspect
ratio less than one

7.5.8.17 Load-Deflection
Analysis of tires deflected by footprint loads generally requires three-di-
mensional models such as the subparametric model described above.
However,, plane-strain modeling can be used for two-dimensional approxi-
mation of the tire's load-deflection behavior.
Plane strain [41] conditions describe stress-strain relationships in struc
tures having one dimension that greatly exceeds the other two dimensions
of the structure. Examples of such structures include tunnels, dams, and
thick cylindrical shells. Such structures may be analyzed by considering a
unit length segment along the large dimension and analyzing the cross sec

o* 20 *B 60 at too
I (MM)

FIGURE 7.43 Calculated deformation pattern due to 165 kPa inflation pressure. Tire aspect
ratio equal to one.
520 MECHANICS OF PNEUMATIC TIRES

s.

ioo -eo -»o MO o 20 «o BO ao 100


Z CMM1

FIGURE 7.44 Calculated deformation pattern due to 165 kPa inflation pressure. Tire aspect
ratio greater than one.

tion. Tires when viewed along the circumferential direction meet the plane
strain criterion and thus may be analyzed by that technique.
Plane strain analysis of an HR78-15 tire was carried out [99] using the
finite element grid of Figure 7.39, as given in ref [99]. The resulting de
flected profile is shown in Figure 7.45; it corresponds to 165 kPa inflation
pressure and 20 mm crown deflection. Calculated responses at selected
levels of inflation pressure and comparison with experimental values are
given in Table 7.4. The plane strain model consistently overestimates the
tire's radial stiffness. This deviation is due to the difference in geometric
stiffnesses associated with circumferential strains in the plane strain model
and the original toroidal tire. The correlation is acceptable because of the
simplicity of the solution technique and the significant reductions in com
putational costs and pre- and post-processing efforts.
A closer correlation was obtained by performing an axisymmetric anal
ysis for the inflation pressure, and using the resulting deformation as ini
tial input to the plane strain analysis. This is a more generalized type of
plane strain than the classical one which assumes zero strain in the third

TABLE 7.4 Responses of H R78- 1 5 tire. Calculations by plane strain analysis

(165 kPa Inflation) (220 kPa Inflation)


Deflection Radial Force (N) Radial Force (N)
(mm) Calc. Meas. Calc. Meas.
5 712 578 670 670
10 1490 1201 1690 1423
15 2335 1868 2802 2269
20 3203 2580 3981 3158
25 4159 3314 5160 4048
TIRE STRESS AND DEFORMATION 521

§8.
^

HO-I20-IOT-BO -60 -40 -20 0 20 4(1 CO BO 100 120 [<0


Z IMM)

FIGURE 7.45 Deflected shape of an HR78-15 tire inflated to 165 kPa and deflected 20 mm
radially

direction. The results are given in Table 7.5. Still closer agreement with
experimental results is expected if the plane strain analysis incorporates
the stresses from an axisymmetric solution as prestressing similar to ther
mal stresses.

7.5.8.18 Future Developments in Analysis


Axisymmetric elements in the isoparametric family have significant ad
vantages over other axisymmetric models; and the plane strain model pro
vides useful characterization of the tire's load-deflection behavior. How
ever, they remain two-dimensional in detail and (a) their stress-strain
formulation by using orthotropic moduli underestimates the material's
stiffness in the belt, (b) orthotropic properties cannot accurately model lo
cations on the tire that have single plies such as the step in the belt. Thus
three-dimensional hexahedrons like the element developed above should
be used for a closer representation of tire behavior and for defining the
state of stress at critical locations in the tire.

TABLE 7.5 Responses ofHR78- 1 5 Tire. Calculations by axisymmetric and plane strain
analysis
(1 65 IcPa Inflation) (220 k Pa Inflation)
Deflection Radial Force (N) Radial Force (N)
(mm) Calc. Meas. Calc. Meas.
5 489 578 534 670
10 1268 1201 1379 1423
15 2024 1868 2482 2269
20 2891 2580 3603 3158
25 3714 3314 4648 4048
522 MECHANICS OF PNEUMATIC TIRES

The results of axisymmetric modeling indicate the need for extreme care
in applying composite theory and calculating equivalent moduli for a
given ply or a group of plies in the tire. The extremely high ratio between
the moduli of the cords and those of the rubbers used in tires makes the
primary function of rubber that of providing a load path between cords
through interply shear. Distinction between plies is best incorporated by
modeling cord-rubber composite plies and the rubber plies in between by
separate elements.
Analysis of tire stresses and deformations due to footprint loadings re
quires further development. The procedure used in ref. [42] follows a cor
rect path by using "hybrid" techniques which allow the use of dis
placement boundary conditions. However, that analysis needs further
refinements before yielding useful results and practical applicability. In
this regard, the "gap element" of ref. [43] shows much promise. Gap ele
ments at the footprint can model a nonlinear foundation behavior with
small stiffnesses before contact and high stiffnesses after the gap has been
closed through contact with the roadway.

7.6 Experimental Techniques


7.6.1 Introduction
Since theoretical methods are not sufficiently developed to allow tire en
gineers to predict the stresses generated in the cord, rubber, and steel com
ponents of a tire under service conditions, a great deal of effort has been
expended developing experimental procedures for this purpose. These
stress and strain measurements are important since they may suggest
changes in tire design which will make more efficient use of construction
materials and/or improve performance, especially with regard to cord and
rubber fatigue phenomena. Additionally, experimental data can be used
to validate analytical predictions in those cases where theory exists. Fi
nally, experiments are often valuable in that they provide guidance in the
development of an accurate mathematical model governing the tire loaded
in some particular manner, i.e., experiments are useful in formulating the
ories.
In experimental mechanics, there are two broad classifications of meth
ods for making strain measurements. These two methods are commonly
referred to as "point" and "whole-field" techniques [44].
In point methods, a strain gage (which could be a mechanical, optical,
or electrical device) is used to measure the average strain in a particular
direction over some given (usually small) gage length. It is apparent that
in many studies, large numbers of strain gages must be used to map a re
gion of interest since such devices function only at a point.
Whole field methods yield data at many points simultaneously. They
are ideally suited for locating regions where large strain gradients or stress
concentrations exist. In the experimental stress analyses of tires, whole
field methods have not been as commonly used as point methods.
The two earliest published articles covering the subject of tire stress
analysis, SchippeFs work [45] of 1923 and Hencky's work [46] of 1935,
make no mention of experimental techniques for measuring strains in
TIRE STRESS AND DEFORMATION 523

tires. In the period 1959-1962, three survey articles appeared which dis
cussed many of the experimental methods known at that time [47-49]. Es
pecially interesting is the work of Kern [47] and of Barson and Gough [49]
who point out the limitation s of several of the techniques especially devel
oped for use with tires.
Methods for measuring strain which are applicable to pneumatic tires
are not particularly different, in principle, from those methods developed
in other areas of experimental stress analysis [50]. However, it has been
necessary to design special transducing systems for tire engineering, since
service strains occur in cord and rubber which are at least an order of
magnitude greater than those which occur in the usual structural materi
als. In addition, the low elastic modulus of cord and rubber, compared to
metals, causes many conventional strain measuring instruments to locally
stiffen the area being measured due to their inherent mechanical rigidity.
This stiffening causes extremely large measurement errors. For this reason
special transducers with low mechanical impedance have been devised to
be elastically compatible with cord and/or rubber.

7.6.2 Measurement of Surface Strain


Many strain measurements made on tires are done by measuring sur
face strains using a variety of techniques. As might be expected, grid
methods play an important role in most of these strain measurements. For
example, Lauterbach and Ames [54] measured the cord strain arising from
the vulcanization and expansion process in an uninflated tire by laying
strips of adhesive tape on the inside of the cured tire from bead to bead.
The tape, when pulled away from the tire, had imprinted on it a replica of
the first ply cord network. When used in conjunction with a mold draw
ing, this replica permitted the cord length changes arising from cure and
expansion to be calculated. For a 4-ply nylon tire, it was established that
cord elongation occurs in the crown region and cord compression in the
bead region. Of course, the sign and magnitude of such cord strains will be
greatly influenced by the parameters of the tire on the building drum rela
tive to the mold geometry, as well as by the thermal shrinkage character
istics of the cord at the tire curing temperature.
Several special transducers have been developed and successfully used
for measuring rubber surface strains in tires. Biderman, et al. [55] reported
data obtained from a transducer which is a modification of the well known
clip gage used for measuring the Poisson contraction during the tensile
testing of metallic specimens. The clip gage has been described in detail by
Perry and Lissner [56]. It is constructed by bonding electrical resistance
strain gages to the upper and lower sides of a U-shaped piece of spring
steel. The legs of the U extend from its backbone, and small pieces of vul
canized rubber are pushed on these pins so that a definite spacing from the
tire surface is achieved when the pins are inserted into the surface of the
article to be tested. This is illustrated in Fig. 7.46. This device shows good
temperature compensation characteristics, and because of the mechanical
attentuation of the signal it can measure strain magnitudes of 20 to 30%,
or larger, depending on the length of the legs. Kern [47], [57-59], has used
this transducer, as well as other types, to obtain strain measurements radi
ally, circumferentially and in the cord direction on the inside and outside
524 MECHANICS OF PNEUMATIC TIRES

surface of passenger car tires. He was the first to study the effect of camber
and cornering on tire strains, and to map the principal strain trajectories in
the contact zone of a deflected tire. Pugin [63] has also used this type of
clip gage to study the effects of crown angle on the sidewall deformation of
bias tires rolling under load.
Liquid metal gages have also been developed to measure the large
strains which occur in rubber. These devices consist of a small rubber tube
or cylindrical cavity in a rubber block filled with mercury or with an in
dium-gallium eutectic. Their principle of operation is similar to that of the
bonded electrical resistance strain gage, i.e. the change in resistance due to
change in geometry is a measure of strain. Hurry and Woolley [66] first
described this type of strain gage but presented no data from actual tire
testing. Gregory, et al. [67] measured inflation strains and some dynamic
strains at various locations on the inside and outside surface of an aircraft
tire using mercury filled capillaries as a sensing element. Similar devices
have been developed to measure strains in human tissue [68]. Janssen and
Walter [69] have used these gages extensively in their experimental studies
on surface strains in bias, belted bias and radial tires. Typical data taken
from their work is shown in Figs. 7.47 and 7.48. These show that at com
parable locations on bias and radial tires the bias design tends to show
more regions of compressive strain, presumably resulting in more frequent
cord compression, than does the radial design.
All of the experimental techniques discussed so far have been point
measurement methods. However there is now considerable interest in dif
ferent types of whole field methods for the measurements of surface
strains in pneumatic tires.
First, it is clear that strains could be measured using surface grids by
placing reference lines on the sidewall of the tire, measuring the distance
between lines before and after loading, and computing the strain as the
change in length divided by the original length. Details of this method are
given by Parks [83]. If the reference lines are in the form of a continuous
rectangular grid pattern, sufficient information is available to determine
the principal strain magnitudes and directions at every point. That is, the
strain can be measured in three directions by noting the change in length
of the sides of the rectangular grid and the change in the length of diago-

STRAIN GAGES

1 r STEEL CLIP

. .. RUBBER WASHER

T— PIN

FIGURE 7.46 Clip gage for measuring large deformations [55]


TIRE STRESS AND DEFORMATION 529
LOCATION 3 LOCATION 10 LOCATION 7
MID-SIDEWALL MID-SIDEWALL CROWN
INNER SURFACE OUTER SURFACE INNER SURFACE

0 10 20 90 0 10 20 30
INFLATION PRESSURE-PSI

FIGURE 7.47 Surface rubber strains measured at the mid-sidewall and crown of a G78-15
belted-bias tire as a function of inflation pressure. Meridional direction O, circumferential
direction V, angular direction Q

nal. Then the principal strains can be calculated from the usual rosette
equation for strain at a point. This method is just a two-dimensional ex
tension of grid methods previously used to measure one-dimensional ex
tension cord strain. Barson and Gough [49] state that a grid method was
their most successful technique for determining surface strains on tires.
The grids were drawn on the rubber using white poster paint and were
transferred to a flat glass surface for measurement using transparent self-
adhesive tape. They report data obtained from statically deflected 5.20-13,
4-ply tires.
Active development is now under way in the application of reflection
photo-elastic techniques to the determination of strains on the surface of
tires. These strains are obtained by detecting changes in the index of re
fraction of light passing through a birefringent material. These materials

LOCATION 3 LOCATION 10 LOCATION 7


MID-SIDEWALL MID-SIDEWALL CROWN
INNER SURFACE OUTER SURFACE INNER SURFACE

10 20 30
INFLATION PRESSURE-PSI

FIGURE 7.48 Surface rubber strains measured at the mid-sidewall and crown of a 20SR-1S
radial tire as a function of inflation pressure. Meridional direction O, circumferential direction
V, angular direction Q
526 MECHANICS OF PNEUMATIC TIRES

are normally epoxy sheets which can either be obtained in preformed con
dition or can be cast and partially polymerized by the experimenter. These
are soft enough to be bonded to the irregular surface of an object such as a
tire, subsequent to spraying of the surface with an aluminum paint which
is used as a reflector. Incident polarized light is then passed through po
larizers and quarter wave plates and through the transparent epoxy sheets
which strain in a manner equal to that of the loaded tire. Light then re
flects from the silver painted surface of the tire and returns through quar
ter wave plates and analyzers to the observer. Since the retardation of the
two polarized beams will be different in the planes of different principal
strain, then these beams emerge from reflection with different velocities
and the resulting retardation is a measure of strain in the body. Such a
photo-elastic effect is analyzed by use of an optical apparatus called a
polariscope. With this device it is possible to determine the differences in
the two principal strains from the isochromatics and the directions of the
principal strains from the isoclinics. The specialized techniques of photo-
elasticity and the solutions to many plane two-dimensional problems of
elasticity are discussed in the treatise of Frocht [84]. An excellent short ex
position of surface photoelastic coating techniques is given by Zandman,
Dally and Redner [65].
Janssen and Walter [87] report data on shear strains near the belt edge
of radial tires using such photo-elastic coatings. These seem to be one of
the most effective ways of studying this phenomenon.
"Moiré" is a French word that describes a silk screen effect and which is
now used to denote an optical phenomenon which can be employed for
surface strain measurement. The Moiré fringe method of experimental
stress analysis is used to measure the relative displacements on a specimen
by the mechanical interference of closely spaced lines. This technique has
been reviewed by Theocaris [88] and is also referred to as mechanical in-
terferometry or the photoscreen method. To use this technique a grid must
be applied to the area of interest and a fringe pattern produced by strain
ing must be measured. The fringes represent contours of equal dis
placement, and strains are calculated from the measured displacements
through the use of the strain-displacement equations. The resulting mea
surement does not depend on a change in resistance (as with an electrical
resistance strain gage) or index of refraction (as with photoelasticity).
Thus, many of the problems inherent in the usual transducing systems (ce
ment creep, stability, zero drift, etc.) can be avoided. Potter [89] has used
the moiré fringe method to measure strains in the range 1-10 percent in a
pressurized neoprene rubber cylinder using two orthogonal sets of parallel
lines with a spacing of 200 lines per inch. He discussed the application of
this technique to tires using long, thin grids with the long axis of the grid
placed radially on the tire sidewall. Browne [90] has used Moiré tech
niques extensively to study fluid film thicknesses in the contact patch of
tires running on flooded surfaces. More recently Moiré methods have been
found useful in studying edge effects in laminated composites, a phenome
non which is difficult to measure quantitatively by other methods. Data on
this has been reported by Pipes [91] and by Oplinger [92].
A technique that has a great deal in common with Moiré is the method
of holographic interferometry based on wavefront reconstruction. A thor
ough description of this technique is given by Haines and Hildebrand [93].
Wavefront reconstruction is a method for recording specific data about a
TIRE STRESS AND DEFORMATION 527

three dimensional object in a medium such as photographic film, and then


reconstructing a three-dimensional image of the object from this record
ing. Holographic fringes, like the Moiré fringes, result when load is ap
plied to a body, and they also represent a locus of points of constant dis
placement. Its advantages compared to other whole-field methods are
increased sensitivity (displacements as small as 10 micro-inches can be
measured) and elimination of the need for mechanical attachments of any
sort to the tire. At the present time holography is much more expensive
than the other methods of experimental stress analysis partly because it re
quires a coherent light source such as a laser. This technique may have
more potential as a nondestructive testing tool than as a quantitative anal
ysis device since non-uniformities in the surface fringe patterns can be
qualitatively interpreted as regions of internal stress concentration in the
tire. This application has been discussed by Grant and Brown [94].
Potts and Corsa [95] have used holographic methods to define the
modal patterns of vibrating tires. This has proven to be a valuable tool for
this purpose since the displacement of the complete tire outer surface may
be obtained on a single photographic image. This allows a clear definition
of the modes involved.

7.63 Measurement of Internal Strain


Measurement of internal strain is considerably more difficult than that
of external strain since the optical and grid methods normally associated
with surface measurements are no longer available to the experimenter.
One type of strain gage which gives promise of yielding strain data on
the internal parts of the tire is the liquid metal gage previously described.
In view of the fact that this gage is extremely compliant, it can be made an
integral part of the tire structure, and it will reflect the extension of the re
gion at which it is located. However these gages are extremely sensitive to
normal pressures and use of them in tires which must be cured under pres
sure is quite difficult.
Another special transducer developed for measuring rubber strains was
first described by Biderman and Pugin [60] and later by Biderman [61],
and is shown in figure 7.49. It consists of a prestressed rubber thread with
a diameter of 1.5 mm. onto which a coil of fine constantan wire has been
wound with a small helical pitch. Compression or elongation of the rubber
thread brings about an increase or decrease in the stress in the wire due to
an increase or decrease in the diameter of the thread. This stressing
changes the electrical resistance of the wire. Prestress of the rubber thread
is necessary in order that the wire remain under mechanical load when the
thread is extended to its maximum elongation. In order to make compres-
sive measurements, the thread must be supported over its entire length on
some sort of base to avoid buckling. For best results the rubber thread
should have a low stiffness and small hysteresis. Calibration and temper
ature compensation of this gage are difficult. Pugin [48], however, recom
mends this transducer for measuring deformations under dynamic condi
tions in preference to the clip gage previously described.
Such rubber-wire gages have been used to measure the circumferential
and radial inflation strains in the proximity of artificial cuts introduced
into the sidewall of agricultural tires [62] and to measure the sidewall
strains developed in radial tires rolling on a drum [63]. Brewer [15] has
528 MECHANICS OF PNEUMATIC TIRES

-RUBBER THREAD

-LEAD WIRE ^FINE WIRE

FIGURE 7.49 Rubber wire transducer for measuring large deformations

used these gages to measure surface strains on a high pressure aircraft tire
in order to compare measured results with calculation. For this purpose,
the gages operated satisfactorily and the results agreed reasonably well
with computation. These gages were at one time manufactured com
mercially by Peekel Laboratorium Voor Electronica, Rotterdam, Holland.
Turner and Ford [100] have developed a particularly simple and effec
tive method for observation and measurement of interply shear strains in
compliant composites, such as typically used in tires. They insert pins
through a two-ply composite of balanced construction with cord angles ±θ
about a centerline, and then observe the tilt, or change in angle, of the pins
as the specimen is loaded. This is illustrated in Fig. 7.50, and it is seen that
interply shear causes the pin to rotate. Fortunately, in compliant com
posites, these rotations are large enough to observe directly, and may be
photographed or measured by other means. Ford and Turner were quite
successful in comparing their measured shear strains with finite element
computations. This technique is valuable in measuring the shear strains in
compliant composites in general, but is particularly valuable in studying

X X
iMXTENDED EXTENDED
FIGURE 7.50 Edge view of a two ply (±6) strip.
TIRE STRESS AND DEFORMATION 529

Epoxy Bond Strain Gages - Lead Wire


TubVto Cord

T«xtil« Cord ^
^Metallic Tub. Uo20in.(typicol)

FIGURE 1.51 Cordforce transducer used by Clark and Dodge [76]

shear in the belt edge region of radial tires, where at the moment it seems
to be the only known technique for working in this important area.

7.6.4 Measurement of Cord Loads


The measurement of cord loads has been one of the most important
fields of activity for the experimental stress analyst in the tire industry. A
large number of methods have been proposed for such measurements.
Among the earliest of these were various techniques for measurement of
cord loads using grid or elongation marks. For example, Furnas [51] used
marks on cords to measure the outer ply cord strains in the sidewall of an
inflated and statically loaded tire, while Loughborough, et al. [52] used the
same technique to measure outer ply cord strain in the sidewall region of
various tires. Most studies establish the fact that in inflated and deflected
bias tires, cords in the contact zone lose the tension initially imposed by in
flation pressure all along their length. Similar techniques were applied by
Buckwalter [53] to measure the strains in the innermost ply of an inflated
and loaded tire, in this case by the use of small knots in the cord as bench
marks. In some cases it has been possible to use X-ray photography to

.030 Dl A.

FIGURE 7.52 Cord force transducer used by Waiter and Hall [80J (Lead wires and strain
gages not shown; dimensions in inches)
530 MECHANICS OF PNEUMATIC TIRES

more clearly define cord strains. Loughborough, et al. [52] used fine steel
wires wrapped tightly around the cords as markers. The markers were put
on the same cord and the change in length between markers due to infla
tion and deflection loads gave a measure of the strain. Weickert [64]
treated one strand of the twisted cord with a metal salt and used the shad
owed cord in X-ray photographs as the marker. He also used small diame
ter steel balls as markers to detect carcass rubber strains between the sec
ond and third ply for a 4-ply nylon tire. All of these methods are limited
by the fact that they can only be used on the static non-rolling tire, and
require careful interpretation of the results.
Miniature force transducers using resistance foil strain gages have been
developed by Patterson [78], Clark and Dodge [76], and Walter [77] which
permit the direct measurement of the cord loads in a tire under operating
conditions. These devices are placed in series with the cord and are em
bedded in the tire during building. In service, they provide a reproducible
and easily monitored electrical signal which is an accurate measure of tire
cord load. These transducers are much smaller than either the clip gage,
the rubber-wire gage, or the liquid metal gages previously discussed.
The force transducer used by Clark and Dodge is shown in figure 7.51;
it is a 0.50 inch long thin-walled beryllium-copper tube with tire cord
bonded through it. Extensive cord force measurements made at the crown,
shoulder, and sidewall in the innermost ply of two- and four-ply tires at
different loads and pressures have been reported using this device [79].
The basic geometry of one of two types of load transducers used by
Walter and Hall [80] is shown in figure 7.52. It is an aluminum alloy billet
of rectangular cross section which averages the force of two adjacent cords
in a tire. This kind of transducer was used to measure cord forces in all
plies of bias and belted bias tires in straight ahead rolling and cornering;
the effect on the cord force pattern of wheel load, inflation pressure, ob-

* jyo _ *>/vi +.

1 C L i
i O
.065
S T
j i
i'C_
!] 7
t 1

1 b 1OI° -
.037 C -* "~
t
\
* .168 " • .tDt *

- ^cr»

FIGURE 7.53 Cord force transducer used by Patterson [78] (lead wires and strain gages not
shown; dimensions hi inches)
TIRE STRESS AND DEFORMATION 531

/TRANSDUCERS -TRANSDUCERS -TRANSDUCERS

\\\

STEP-IN STEP-OUT BEAD REINFORCE


"2-0"Tie-ln

FIGURE 7.54 The three constructions for wrapping fabric around the bead of two ply tires.

stack impact, tire speed, rim width, and tire-road interface was also stud
ied. Significant ply-to-ply cord force variations were detected, and the
cord force patterns observed in the first ply as the transducer passed
through the footprint were nearly mirror images of those observed in the
second ply at the same location. The cord tension developed at various lo
cations in an inflated and rotating, but otherwise unloaded, two-ply poly
ester tire was measured with this device up to angular velocities of 1600
rpm—about 120 mph [81]. Patterson [78] used a force transducer 0.460
inch long of relatively complex geometry to measure the cord loads which
occur during shaping, cure, and postinflation of a 7.50-14 two-ply tire as
shown in fig. 7.53. These are the only published data treating this subject.
Similar cord load measurements during shaping and curing of belted-bias
and radial tires have been reported by Janssen and Walter [87].
Some measure of the usefulness of these miniature force transducers is
afforded in the investigation of cord loads in the ply turn-up region, i.e.,
near the bead. In this area, changes in tire stiffness and the closeness of the
rim render theoretical stress analysis difficult if not impossible. Even sur
face rubber strains on the inside of the tire would be difficult to obtain in
this otherwise inaccessible region. In the turn-up area, design variations
exist among manufacturers because, for tires with two body plies, three
different methods exist for wrapping fabric around the beads: a 2-0 tie-in
with a step-in construction, a 2-0 tie-in with a step-out construction, and a
1-1 tie-in with a bead reinforce construction. These different constructions
are shown in figure 7.54. Cord force data given by Walter [82] obtained in
the middle of the step and at the edge of the reinforce are shown in figure
7.55 for otherwise identically constructed G78-15 belted bias tires with
two polyester body plies and two fiberglass belts at rated load (1380 lb.)
and pressure (24 psi) in straight ahead rolling at low speed (2 mph). In
each case, the magnitude of the initially imposed inflation tension and the
peak-to-peak cord force values are consistent with the type of turn-up de
sign.

7.6.5 Measurement of Bead Forces


Only recently have serious efforts been made to measure bead forces in
pneumatic tires. Stiebel [97] has recently developed a very clever tech-
532 MECHANICS OF PNEUMATIC TIRES

"2-0"Tie-1n "2-0"Tie-ln Vl" Tie-In


STEP-IN STEP-OUT BEAD -REINFORCE

•±I808 Oe ±180° ±180° 0° ±180° ±180° 0°


±180°
ANGULAR LOCATION OF TRANSDUCER RELATIVE TO CENTER OF
CONTACT ZONE (0°)

FIGURE 1.55 Cordforces measured at rated load and pressure in each ply in the turn-up
region for G78-15 belted bias tires [82]. (See Fig. 7.53 for transducer locations).

nique for the measurement of strains in bead wire. This was accomplished
by using unbonded strain gage wire wrapped in the form of a helix around
an individual bead wire. If the strain gage wire is properly insulated then
it can act as a strain gage, the double helix wrap giving compensation both

BIFILAR WOUND STRAIN GAGE

FIGURE 7. 56 Bifilar wound bead wire strain gage


TIRE STRESS AND DEFORMATION 533

Ib s
200- G78-15

m
100-
CO
O"~
0—
•• * " JV \
V****Xl«*XX*>^XtfxxKXXx)c ,,,,,, x>>£<
COMPRES ION

100-

•A2?i A1/f " ^^


200-
It s

012345678 9101112131415
CURE TIME IN MINUTES

FIGURE 7.57 Bead loads measured during curing cycle on individual strands for a bias ply
tire.

for bending and torsion. The configuration of the gage used by Stiebel is
shown in Fig. 7.56, while typical bead loads measured during curing are
given in Figs. 7.57 and 7.58.
Measurements of contact force between tire and bead are difficult to
make, but with the advent of relatively small pressure transducers such
measurements are now possible. A series of such measurements on a vari
ety of tires was reported by Walter and Kiminecz [98]. They measured
these pressures using a miniature transducer as the load sensor. Typical
values which they report for a JR78-15 tire on a 15 × 6 rim during straight
line rolling are given in Fig. 7.59.

7.7 Perspective
The pneumatic tire is one of the most challenging structures that an ex
perimentalist or analyst may face. This is due to several factors, including:
(1) Complications in the composite material caused by (a) in-
compressibility of the rubber, which requires special stress-strain re
lations and influences the numerical condition of the constitutive
equations, (b) the twisted construction of the cords makes them an i-
sotropic and yields different properties in tension and compression,
and (c) the ratio between the modulus of elasticity of the cord and
that of the rubber is far in excess of corresponding ratios found in
534 MECHANICS OF PNEUMATIC TIRES

200 GR78-15

100-

012345678 9101112131415
CURE TIME IN MINUTES

FIGURE 7.58 Bead loads measured during curing cycle on individual strandsfor a radial ply
lire.

most composites; this leads to severe stress gradients at their inter


face and puts the accurate prediction of properties of cord-rubber
composites beyond conventional composite theories.
(2) The formation pattern of a tire deflected by footprint loads involves
sharp changes in geometry; flattening a doubly curved segment of
the toroid yields sharp curvature changes at the footprint bounda
ries and results in complicated stress fields.
(3) Due to the nature of tire materials, conventional resistance strain
gages are not directly usable to measure most tire strains. These
gages are normally the main tool used in experimental stress analy
sis. Hence, special techniques must be developed and used for tire
strain measurement.
(4) Many tire stresses of the greatest practical interest are extremely dif
ficult to measure because they occur in very local areas, and good
techniques for such measurements do not exist. Examples of these
are stresses in the interface between cord and rubber, at the adhe
sive layer, and stresses in the belt edge area of radial tires. (5) The
(5) The magnitude and distribution of many tire loadings have not yet
been fully defined.
For a long time these difficulties discouraged the expenditure of signifi
cant efforts on direct computational analysis of tires. Simplified models
and semi-empirical formulations were used as substitute procedures, and
provided much of our early understanding of trends in tire behavior.
TIRE STRESS AND DEFORMATION 535

100 -

300-

200

±180- -120' -60- 0' +60 +120* ±180'


TRANSDUCER ANGULAR LOCATION

FIGURE 7.59 Contact pressures between bead area and rim of a typical tire.

However, in recent years two factors have increased the emphasis on tire
structural analysis. These are (a) a new outlook in design of tires and other
vehicle components demanding the analysis of these products from first
principles, and (b) the availability of better analytical and numerical
methods and larger and faster computers. The analytical techniques dis
cussed in this chapter help build a foundation for attacking this challeng
ing task.
The progress made in the finite element field deserves special attention.
The problems of calculating tire deformations due to inflation pressure
and centrifugal forces, calculation of cord forces, and predicting the radial
spring rates of tires have been solved using these techniques. Analysis of
tire stresses at critical locations such as belt edges and other ply discontin
uities requires refined analyses. Single plies are anisotropic, thus necessi
tating the use of "twist" element models which allow circumferential (i.e.,
twist) displacements due to axisymmetric loads (e.g. tire inflation). Future
investigations may also show that more detailed models, representing cord
and rubber separately, are needed for accurate analysis of stresses and de
formations within and between tire plies.
While some progress has been made in developing specialized trans
ducers and other techniques for tire experimental stress anlaysis, there still
exists a great need for more refined methods, particularly methods allow
ing mapping of stress fields in very small regions, of the order of 1 mm, in
the interior of a tire structure.
The activitites currently underway are expected to shed further light on
the degree of refinement needed for obtaining results at desired detail lev
els. This chapter provides fundamentals which will aid in understanding
present and future techniques.
536 MECHANICS OF PNEUMATIC TIRES

7.8 Principal Notation


A matrix relating membrane forces to strains
B bending-stretching matrix
b matrix of undetermined coefficients b,
C stress-strain matrix
c tire parameter denned in eq. (7.12)
D matrix relating tire bending moments to tire curvatures
F matrix relating displacement matrix <j> to the undetermined
coefficients b,,
F, matrix F evaluated at element nodes
/ polynomial function for subparametric element
g matrix relating strains to coefficients b
hK distance from laminate midplane to tCh ply
J Jacobian matrix relating derivatives
k stiffness
k element stiffness matrix
Kt, Kg. K+g changes in curvature and torsion
Mt, M, bending moments in the tire
M+, twisting moment in the tire
N total number of cords in the tire
NV Ne membrane forces per unit length
N+g membrane shear force per unit length
p inflation pressure
P applied load
</„, </„ external loads on the tire
Q+ transverse shear force
Q matrix relating stresses to strains
r radial coordinate
rt meridional radius of curvature
r, circumferential radius of curvature
rc radial coordinate of tire crown
rw radial coordinate of the widest section of the tire
s matrix related to element stiffness matrix k
/ radial ply designation
T cord tension force
u tangential displacement
u, circumferential displacement
u, radial displacement
u, axial displacement
u, displacements of corner nodes
K- normal displacement
z axial coordinate
a tire parameter defined in eq. (7.10)
ft cord angle in bias tires
ft. cord angle at tire crown
•y tire parameter defined in eq. (7.13)
f> displacement
cj. **> y#> strains at tire neutral axis
CM, e" I radial, axial, circumferential, and shear strains
| strains in the nth term of Fourier solution
TIRE STRESS AND DEFORMATION 537

6 circumferential coordinate or angle


X parameter defining meridional length of a shell
fi tire parameter denned in eq. (7.1 1)
>W> "^ Poisson's ratios
£, TJ, f curvilinear axes
o?X,
<j,°, T°n initial element stresses
^ tire variable denned in eq. (7.20)
u* rotations
<£ meridional angle
<> matrix of u displacements and slopes
9,. matrix <?> evaluated at element nodes

7.9 Acknowledgement
The authors wish to thank all the scientists whose work was referenced
in this chapter, and specifically Dr. J. Padovan for supplying ref. 22 and
Dr. J. D. Walter for a number of helpful suggestions. We also thank Mr.
F. S. Conant for reading the manuscript and offering valuable suggestions,
Mrs. M. I. Zeigler for typing the manuscript, Mrs. P. J. Lindeman for as
sistance in typing and preparing the illustrations, and the Firestone Tire &
Rubber Company for permission to publish this work.

REFERENCES
[1] Alliger, G., "The Cordless Tyre—Its Performance and Market Potential," Inter
national Rubber Conference, Brighton, May 1972.
[2] Ridha, R. A., Analysis for Tire Mold Design, Tire Science and Technology, 1, 195-
210, 1974.
[3] Clark, S. K. (Editor), Mechanics of Pneumatic Tires, National Bureau of Standards
Monograph 122, Washington, D.C., 1971.
[4] Frank, F., and Hofferberth, W., Mechanics of the Pneumatic Tire, Rubber Chemistry
and Technology, 40, 271-322, 1967.
[5] Walter, J. D., Tire Stress and Deformation, (In Mechanics of Pneumatic Tires, S. K.
Clark, Ed., NBS Monograph 122, Washington D.C., 1971).
|6] Harkleroad, W. I., Basic Principles of Hose Design, Rubber Chemistry and Tech
nology, 42, 666-674, 1969.
[7] Faupel, J. H., Engineering Design, pp. 312-317 (J. Wiley & Sons, New York, 1964).
[8] Timoshenko, S., and Woinowsky-Kreiger, S., Theory of Plates and Shells, Second Edi
tion, (McGraw-Hill Book Co., New York, 1959).
[9] Midgely, T., Tire Construction, U.S. Patent 1,802,088, filed Oct. 7, 1920.
[10] Gough, V. E., Cord Path in Tyres, Transactions of the Institution of the Rubber Indus
try, 40, T20-T57, 1964.
[11] Robecchi, E., and Amici, L., Mechanics of the Pneumatic Tire, Tire Science and Tech
nology, 1,290-345, 1973.
[12] Tielking, J. T. and Feng, W. W., The Application of the Maximum Potential Energy
Principle to Nonlinear Axisymmetric Membrane Problems, Journal of Applied Me
chanics, 73, 491-496, (1974).
[13] Feng, W. W., Tielking, J. T., and Huang, P., The Inflation and Contact Constraint of a
Rectangular Mooney Membrane, Journal of Applied Mechanics, 73, 979-983, 1974.
[14] DeEskinazi, J., Werner, S., and Yang, T. Y., Contact of an Inflated Toroidal Mem
brane with a Flat Surface as an Approach to the Tire Deflection Problem, Tire Sci
ence and Technology, 3, 43-61, 1975.
[15] Brewer, H. K., Stresses and Deformations in Multi-Ply Aircraft Tires Subject to Infla
tion Pressure Loadings, Technical Report AFFDL-TR-70-62, Wright-Patterson Air
Force Base, Ohio, June 1970.
538 MECHANICS OF PNEUMATIC TIRES

[16] Kalmns. A., Analysis of Shells of Revolution Subjected to Symmetrical and Non-Sym
metrical Loads, Journal of Applied Mechanics, 31, 223, 1964.
[17] Hrennikoff, A., Solution of Problems of Elasticity by the Framework Method, Journal
of Applied Mechanics, ASME, 169-175, 1941.
[18] Hall, A. S., and Woodhead, R. W., Frame Analysis, (John Wiley 4 Sons, New York,
1961).
[19] Wilson, E. I... Structural Analysis of Axisymmetric Solids, AIAA Journal, 3, 2269-
2274, 1965.
[20] Dunham, R. S., and Nickel), R. E., Finite Element Analysis of Axisymmetric Solids
with Arbitrary Loadings, Report No. 67-6, Department of Civil Engineering, Uni
versity of California, Berkeley, California, June 1967.
[21] Ridha, R. A., Finite Element Stress Analysis of Automotive Wheels, Trans. Soc. Auto.
Eng. V.85, 1976, p. 301 (SAE Paper No. 760085).
[22] Kaga, H , Okamoto, K., and Toxawa, Y., Internal Stress Analysis of the Tire Under
Vertical Loads Using Finite Element Method, Tire Science and Technology, 5, p.
102-118, 1977.
[23] Kelsey, S., Lee, K. N., and Mak, C. K. K., The Condition of Some Finite Element Co
efficient Matrices, Study No. 5, Computer- Aided Engineering, University of Water
loo, Ontario, Canada, May 1971, pp 267-283.
[24] Oliveira, E. R. A., Optimization of Finite Element Solutions, Third Conference on
Matrix Methods in Structural Mechanics, Wright-Patterson Air Force Base, Ohio,
October 1971.
[25] Zorowski, C. E., and Dunn, S. E., A Mathematical Model for the Pneumatic Tire,
North Carolina State University, Raleigh, June 1970.
[26] Walter, J. D., Centrifugal Effects in Inflated, Rotating Bias Ply Tires, Textile Research
Journal, 46, 1-7, 1970.
[27] Patel, H. P., Mathematical Analysis of Statically Loaded Penumatic Tires, Ph.D
Thesis, North Carolina State University, Raleigh, 1975.
[28] Zorowski, C. F., and O'Neil, E. W., Measurement of Pneumatic Tire Contact Pressure
for Static Loading, Report to the Office of Vehicle Systems Research, North Caro
lina State University, Raleigh, 1969.
[29] Turner, M. J., Clough, R. W., Martin, H. C., and Topp, L. J., Stiffness and Deflection
Analysis of Complex Structures, Journal of Aeronautical Sciences, 23, 805-823,
1956.
[30] Gallagher, R. H., Padlag, J., and Bijlaard, P. P., Stress Analysis of Heated Complex
Shapes, Journal of Aerospace Science, 32, 700-717, 1962.
[31] Melosh, R. J., Structural Analysis of Solids, Journal of the Structural Division, ASCE,
89, No. ST4, 205-224, 1963.
[32] Clough, R. W., Comparison of Three- Dimensional Finite Elements, Proc. Symposium
on Application of Finite Element Methods in Civil Engineering, Vanderbilt Univ.,
November 1969, pp 1-26.
[33] Ergatoudis, I., Irons, B. M . and Zienkiewicz, O. C., Curved, Isoparametric, Quadrilat
eral Elements for Finite Element Analysis, International Journal of Solids and
Structures, 4,31-42, 1968.
[34] Zienkiewicz, O. C., Irons, B. M., Ergatoudis, J., Ahmed, S., and Scott, F. C., Isopara
metric and Associated Element Families for Two- and Three-Dimensional Analysis,
Finite Element Methods in Stress Analysis, TAPIR, Technical University of Nor
way, Trondheim, 1969, Chapter 13.
[35] Ridha, R. A., Kelsey, S., Cervelli, R. V., and Cadoret, J. E., Proposal to Supply a
Computer Program for the Finite Element Analysis of Aircraft Wheel Assemblies,
Report No. VP-1 122, The Bendix Corporation, South Bend, Indiana, January 1972.
[36] Chacour, S., "DATUNA", A Three-Dimensional Finite Element Program used in the
Analysis of Turbomachinery, ASME Paper No. 71-WA/FE-29, Winter Annual
Meeting, Washington, D.C., December 1971.
[37] Adelman, H. M., Catherines, D. S., and Walton, W. C., Accuracy of Model Stress Cal
culations by the Finite Element Method, AIAA Journal, 8 462-468, 1970.
[38] Fung, Y. C., Foundations of Solid Mechanics, Prentice-Hall Inc., Englewood, N. J.,
1965, p 157.
[39] Irons, B. M., Numerical Integration Applied to Finite Element Methods, unpublished
paper (written around 1966).
[40] Zienkiewicz, O. C., and Nayak, G. C., A General Approach to Problems of Large De
formation and Plasticity Using Iso- Parametric Elements, Third Conference on Ma
trix Methods in Structural Mechanics, Wright Patterson AFB, Ohio, October 1971.
[41] Timoshenko, S., and Goodier, J. N., Theory of Elasticity, (McGraw-Hill Book Co-
New York, 1961).
[42] Deak, A. L., and Atluri, S., The Stress Analysis of Loaded Aircraft Tires, Interim Re
port for AFFDL, Mathematical Sciences Northwest. Inc., 1973.
TIRE STRESS AND DEFORMATION 539

[43] Quarterly News Letter, Marc Analysis Research Corporation, Providence, Rhode Is
land, Third Quarter, 1976.
[44] Dove, R. C. and Adams, P. H., Experimental Stress Analysis and Motion Measure
ment, p. 18 (Merrill, Columbus, Ohio, 1964).
[45] Schippel, H. F., Fabric stresses in pneumatic tires, Industrial and Engineering Chemis
try 15, 1121-1131 (1923).
[46] Hencky, H., Stresses in rubber tires, Mechanical Engineering 57, 149-153 (1935).
[47] Kern, W. F., Strain measurements on tires by means of strain gauges, Revue Generate
du Caoutchouc 36, 1347-1365 (1959); Kautschuk and Gummi 13, WT59-WT68
(1960).
[48] Pugin, V. A., Electrical strain gages for measuring large deformations, Soviet Rubber
Technology 19 (1), 23-26 (1960).
[49] Barson, C. W. and dough, V. E., Measurement of strain of materials of large exten
sibility, British Journal of Applied Physics 13, 168-170 (1962)
[50] Hetenyi, M. (ed.), Handbook of Experimental Stress Analysis, (Wiley, New York,
1950).
[51] Furnas, A. R., Firestone Tire & Rubber Co. Research Report, (1933) (unpublished).
[52] Loughborough, D. I,., Davies, J. M. and Monfore, G. E., The measurements of strains
in tires, Canadian Journal of Research 28 (F), 490-501 (1950); Revue Generate du
Caoutchouc 29, 712-716 (1952).
[53] Buckwalter, H. M., United States Rubber Co. Research Report, (1932) (unpublished);
ASTM Committee D 13 Meeting, Providence, Rhode Island, (1933).
[54] Lauterbach, H. G. and Ames, W. F., Cord stresses in inflated tires, Textile Research
Journal 29, 890-900 (1959).
[55] Biderman, V. L., Drozhzhin, P. K., Pugin, V. A. and Shchaveleva, V. F., Experimental
Investigation of Deformations in Elements of Pneumatic Tires (Russian), Transac
tions Tire Research Institute (NIIShP) 3, pp. 5-15 (State Scientific and Technical
Publishing House, Moscow, 1957).
[56] Perry, C. C. and Lissner, H. R., The Strain Gage Primer, p. 275 (McGraw-Hill, New
York, 1962).
[57] Kern, W. F., Uber die Cordbeanspruchung in Reifen, Faserforschung und Textil-tech-
nik 11, 401-408 (1960).
[58] Kern, W. F., Uber Verformungsmessungen an Kraftfahrzeugreifen mittels spezieller
Dehnungsgeber, Automobiltechnische Zeitschrift 63, 33-41 (1961).
[59] Kern, W. F., Mather, K. and Nippold, H., Uber die Auswertung von Verformungs-
Messungen an Reifen, Kautschuk und Gummi Kunststofle 16, 619-627 (1963).
[60] Biderman, V. L. and Pugin, V. A., Wire Strain Gauge for Measurements of Large De
formations (Russian), Zavodskaya Laboratoriya 24 (7), 874-875 (1958).
[61] Biderman, V. L., Deformation of tyre elements during rolling, Revue Generate du
Caoutchouc 36, 1366-1371 (1959).
[62] Dzhafarov, I. D.. Deformation in the damage zone of a tyre, Soviet Rubber Tech
nology 26 (3), 38-39 (1967).
[63] Biderman, V. L., Pugin, V. A. and Fil'ko, G. S., Deformation and stresses in the cover
rubber of the sidewall of a radial tyre, Soviet Rubber Technology 24 (7), 18-20
(1965).
[64] Weickert, B., Methods for measuring static strains in automobile tires, Textile Re
search Journal 32, 705-710 (1962).
[65] Zandman, F., S. Redner and J. W. Dally, "Photoelastic Coatings," Iowa State Univer-
isty Press 1977.
[66] Hurry, J. A. and Woolley, R. P., A new high range strain gage, Rubber Age 73, 799-
800 (1953).
[67] Gregory, R. K., Rastrelli, L. U. and Minor, J. E., Tire Structural Design Improvement,
Contract AF33(657)- 12861 with Air Force Systems Command, Southwest Research
Institute, 1968 (ASD-TR-68-12).
[68] Gibbons, G. E., Strandness, D. E. and Bell, J. W., Improvements in design of the mer
cury strain gauge plethysmograph, Surgery, Gynecology, & Obstetrics 115, 679-682
(1963).
[69] Janssen, M. L. and Walter, J. D., "Rubber Strain Measurements in Bias, Belted-Bias
and Radial Ply Tires", Jour. Coated Fibrous Materials", v. 1, Oct. 1971, p. 102.
[70] Alekseev, P. I., Stressing of the bead rings of a tyre under various conditions of defor
mation, Soviet Rubber Technology 19 (2), 18-22 (1960).
[71] Pugin, V. A., Stresses in the bead rings of tyres, Soviet Rubber Technology 21 (10), 8-
12 (1962).
72] Forster, M. J., Firestone Tire & Rubber Co. Research Report, (1952) (unpublished).
[72]
73] Kern, W. F., Observations on the relation between laboratory and test stand measure
[73]
ments of tire treads and their behavior on the road, Rubber Chemistry and Tech
nology 29, 806-828 (1956).
540 MECHANICS OF PNEUMATIC TIRES

[74] Biderman, V. L., et al, Basic Theory of Pneumatic Tires (Russian), Automobile Tires
(Construction, Design, Testing, and Usage), pp. 46-171 (State Scientific and Techni
cal Publishing House, Moscow, 1963).
[75] Biderman, V. L., Critical Speed of a Rotating Pneumatic Tire—Basic Experimental
Data (Russian), Raschet na Prochnost No. 7, Part 3, pp. 324-349 (Moscow, 1961).
[76] Clark, S. K. and Dodge, R. N., Development of a textile cord load transducer, ORA
Report 01 193-1-T, (University of Michigan, 1968).
[77] Walter, J. D., A Tirecord Tension Transducer, Textile Research Journal 39, 191-1%
(1969).
[78] Patterson, R. G., The Measurement of Cord Tensions in Tires, Rubber Chemistry and
Technology 42, 812-822 (1969).
[79] Clark, S. K. and Dodge, R. N., A Load Transducer for Tire Cord, SAE Paper No.
690521, (1969).
[80] Walter, J. D. and Hall, G. L., Cord load characteristics in bias and belted-bias tires,
SAE Paper No. 690522 (1969).
[81] Walter, J. D., Centrifugal effects in inflated, rotating bias ply tires. Textile Research
Journal, 40, 1-7 (1970).
[82 Walter, J. D., Firestone Tire & Rubber Co. Research Report, (1969) (unpublished).
S3 Parks, V. J., The Grid Method, Experimental Mechanics 9 (7), 27N-33N (1969).
84 Frocht, M. M., Photoelasticity, (Wiley, New York, 1941 (Vol. I), 1948 (Vol. II)).
85 Angioletti, A., Eccher, S., Polvara, O. and Zerbini, V., Rubber birefringence and
photoelasticity, Rubber Chemistry and Technology 38, 1 1 15-1 163 (1965).
[86] Oppel, G., Konstruktionsprufung von Gummifederungen fur Motorfahrzeuge mit
dem Eintrevertahren und mit Spannungsoptik, Automobiltechnnche Zeitschrift 51,
77-84 (1949).
[87] Janssen, M. L., and Walter, J. D., "Stresses and Strains in Tires", Tire Sci. and Tech.,
v. 3, n. 2, May 1975.
[88] Theocaris, P. S., Moiré Fringes: A Powerful Measuring Device, Applied Mechanics
Reviews 15, 333-339 (1962).
[89] Potter, J. M., Measurement of Strain on Curved Surfaces Using the Moiré Method,
M.S. Thesis, University of Illinois, (1968).
[90] Browne, A. L., "Tire Deformation During Dynamic Hydroplaning", Tire Sci. and
Tech., v. 3, n. 1, Feb. 1975.
[91] Pipes, R. B. and Daniel, I. M., "Moiré Analysis of the Interlaminar Shear Edge Effect
in Laminated Composites", J. of Composite Materials, 5, 255 (1971).
[92] Oplinger, D. W , Parker, B. S. and Chiang, F. P., "Edge Effect Studies in Fiber-Rein
forced Laminates", Experimental Mechanics 14, 347 (1974).
[93] Haines, K. A. and Hildebrand, B. P., Surface-Deformation Measurement Using the
Wavefront Reconstruction Technique, Applied Optics 5, 595-602 (1966).
[94] Grant, R. M. and Brown, G. M., Holographic Nondestructive Testing (HNDT) in the
Automotive Industry, SAE Paper No. 690051, (1969).
[95] Potts, G. R. and Corsa, T. T., "Tire Vibration Studies: The State of the Art", Tire Sci.
and Tech., v. 3, n. 3, Aug. 1975, p. 135.
[96] Nicholson, D. W., "On the Inflated Tire Profile", Tire Science and Technology, v. 4, n.
3 August 1976 p. 169.
[97] Stiebel, A., "Instrumentation Techniques for Measuring Bead Stresses in Tires" Soc.
Auto. Engr's., Paper No. 770874.
[98] Cook, R. D., "Concepts and Applications of Finite Element Analysis," John Wiley &
Sons, New York.
[99] Ridha. R. A., "Analysis of Pressurized Composite Shells Deflected Against Rigid Flat
Surface," Preceding Symposium on Applications of Computer Methods in Engi
neering, University of Southern California, Los Angeles, Vol. II, Aug. 1977, pp.
1075-1083 (L. C. Wellford, Ed.).
[100] Turner, J. L. and Ford, J. L., "Interply Behavior Exhibited in Compliant Filamentary
Composite Laminates," Paper presented at Soc. Exp. Stress Analysis, Fall Meeting,
October 11-13, 1977, Philadelphia, Pennsylvania.
Chapter 8
MEASUREMENT OF TIRE
PROPERTIES
H. van Eldik Thieme1, A. J. Dijks1, Stephen Bobo2

8.1 Introduction 542


8.1.1 Nomenclature 543
8.1.2 List of Symbols 544
8.2 Straight Line Rolling Experiments 546
8.2. 1 Introduction to Total Force and Moment Measure
ments 546
8.2.2 Load Deflection Relationships 559
8.2.3 Effective Rolling Radius 577
8.2.4 Rolling Resistance 584
8.2.5 Braking and Traction 599
8.2.6 Tire Nonuniformities 612
by A. Dijks
8.2.7 Tire Flaws and Separations 636
by Stephen Bobo
8.3 Cornering and Camber Experiments 658
8.3. 1 Introduction 658
8.3.2 Cornering Experiments 664
8.3.3 Camber and Cornering 688
8.3.4 The Influence of Braking and Traction on Corner
ing 695
8.3.5 Difficulties in Measuring Forces and Moments 702

1 Dep'l of Mechanical Engineering, Delft Technical University. Delft, Netherlands.


2 US Dep'l. of Transportation, Transportation Systems Center, Cambridge, Mass.

541
542 MECHANICS OF PNEUMATIC TIRES

MEASUREMENT OF TIRE PROPERTIES

8.1. Introduction
The combination of road, tire, vehicle, and driver forms one entity. The
mechanical characteristics of the tire in contact with the road must com
bine with the mechanics of the vehicle to help in producing operational
characteristics of the tire-vehicle system which are satisfactory to the
driver.
"The complexity of the structure and behavior of the tire are such that
no complete and satisfactory theory has yet been propounded. The char
acteristics of the tire still present a challenge to the natural philosopher to
devise a theory which shall coordinate the vast mass of empirical data and
give some guidance to the manufacturer and user. This is an inviting field
for the application of mathematics to the physical world."
In this way Temple formulated the situation of more than one decade
ago (Endeavour, October 1956). Since then, in numerous institutes and
laboratories, the work of the earlier investigators has been continued.
Considerable progress in the development of tire mechanics during the
last decade has led to a better understanding of tire behavior. Owing to the
infinite complexity of the pneumatic tire and its interaction with the road
it does not appear at present, despite the progress made, that Temple's
view will be altered in the foreseeable future. Thanks to new and more re
fined experimental techniques becoming increasingly available, and to the
introduction of the electronic computer, the goal of formulating more real
istic mathematical models based on better insight and leading to more re
liable prediction of tire performance may be achieved.
The authors of this chapter do not claim to have supplied a picture of
tire behavior which covers all knowledge achieved hitherto. A selection of
studies has been made in order to provide the engineer and the student
with background material necessary for the investigation and the under
standing of tire and vehicle functional performance.
From the point of view of the engineer and the applied mathematician
the mechanical behavior of the tire must be systematically investigated in
terms of its reaction to various kinds of input related to vehicle motions
and road parameters.
With reference to the role of a tire it is convenient to distinguish be
tween symmetric and anti-symmetric modes of performance. First, the tire
supports the vertical axle load and transmits longitudinal braking or driv
ing forces. Second, the tire is called upon to supply the lateral cornering
and camber forces which are necessary for the directional control of the
vehicle. The content of this chapter has been subdivided according to
these categories.
Many of the investigations discussed in this chapter have been carried
out at the Vehicle Research Laboratory of the University of Technology,
Delft, Holland.
MEASUREMENT OF TIRE PROPERTIES 543

The authors wish to express their appreciation to the members of the


staff of this laboratory: Especially to A. Dijks for contributing to parts of
the text as well as reviewing pans during its preparation, J. van den Berg
and J. A. Zwaan; E. G. M. J. de Vries and his electronic measuring depart
ment; D. A. Timan, J. H. M. Rooney and P. J. Jillesma for their numerous
tire experiments, as well as to H. M. Snijders and his workshop for the as
sistance in manufacturing various instruments and apparatus.

8.1.1. Nomenclature
For both the experimental and theoretical investigations of tire behavior
described in this chapter, we have attempted to use a uniform system of
notation. As a rule, the meaning of symbols has been explained in the text.
For this reason, only a list of the most important symbols will be given be
low. The choice of symbols has been inspired by the list which has been
proposed by a SAE committee (cf. ref. [3] indicated at the end of sec. 8.3).
A number of changes and additions appeared to be necessary in order to
obtain a more or less systematic and usable system of symbols adjusted to
the specific subjects of this chapter.
Constant quantities describing construction, configuration and proper
ties of the real tire or of the theoretical model are defined in such a way
that they become positive. In most cases, the positive sense of variable
quantities are chosen in accordance with the (C, x, y, z) system of axes
shown in figure 8.1.1. The origin C, defined as contact center, is the point
of intersection of the road-plane, the wheel center-plane and the plane
which is situated normal to the road-plane and which passes the wheel
axis. The jc-axis points forward and forms the intersection of the wheel
center-plane and road-plane. The z-axis points downward and is directed
perpendicular to the road-plane. Consequently, the j>-axis is the per
pendicular projection of the wheel axis onto the road. In the same figure
8.1.1, the positive directions of forces and moments acting from road to
tire have been indicated, as well as the positive senses of the variables
which describe the deviations of the position and the motion of the wheel
center-plane with respect to the rectilinear steady state motion of the
wheel center-plane, which in that case coincides with the (x, z) plane of
the coordinate system (0, x, y, z) fixed to the road with the f-axis directed
vertically. As in figure 8.1.1, the road-plane has in most cases been consid
ered as a smooth horizontal surface.
In some cases an alternative definition of positive sense has been felt to
be preferable. In order to work with positive quantities, the tire normal
load has been defined as FN(= W) = -FZ. Similarly, the quantity Fr = - Fx
has been introduced, denoting the rolling resistance force during free roll
ing, i.e., at constant forward velocity and without traction or braking
torques. Also, the sense of the speed of rotation SI of the forward rolling
wheel has been defined as positive. Sometimes, the absolute values of the
longitudinal force Fx have been considered. They are designated as the
braking force FB(= —F,) and the traction force FT(=Fx).
The lateral force acting from road to tire, F^, has been provided with an
additional subscript α or γ in cases when it has been felt necessary to ex
press whether side slip or camber causes the lateral force.
544 MECHANICS OF PNEUMATIC TIRES

normal to road-plane

wheel-centre-plane

wheel-axis

steady state (stationary )


rectilinear free rolling

FIGURE 8.1.1 Nomenclature and coordinate system for a wheel on a plane surface.

8.1.2 List of Symbols


L, F and T denote length, force and time units respectively.
RAD denotes radians.
a half length of contact area (L)
b half width of contact area (L)
c,f, foundation stiffness per unit length in tangential (/), lateral
(c) and radial direction (r) respectively (F/L2)
stiffness of tread rubber per unit area in longitudinal (,v) and
lateral direction (y) respectively (F/L3)
c, tensile tread band (carcass) stiffness per unit length (F)
Cf,,y carcass stiffness in contact region in x and y directions re
spectively (F/L)
CFa (=dFy/da at a = 0) cornering stiffness (cornering rate) (F/
RAD)
(= dFJdy at y = a = 0) camber rate (cf. eq (9.5.61)) (F)
(=—dMJda at a = 0) cornering stiffness (aligning rate) (FL/
RAD)
(- dMr/df at S = S0) rolling resistance coefficient (F)
MEASUREMENT OF TIRE PROPERTIES 545

C, (= - dFJdx, or dFJdr^ at F, = 0) longitudinal or tangential


stiffness in contact region of non-rolling tire (F/L)
C, (= - dFf/dya at Fy = 0) lateral tire stiffness in contact region
of non-rolling tire (F/L)
C, (= dW/df = dW/dza at 8 = da or za = 0) normal tire stiffness
(F/L)
C, (- dFJdx at K = 0) longitudinal slip stiffness (F)
Q (= - dM2/dfy at M, = 0) torsional stiffness about vertical
axis of non-rolling tire (FL/RAD)
El flexural rigidity of tread band (FL1)
FB T braking and traction force respectively (F)
FH (= W = - Fc) tire normal load (F)
F,JJ longitudinal, cornering (= lateral) and normal force acting
from road to tire (cf. fig. 8.1.1) (F)
G(x, Q Green's function
Ix polar moment of inertia of wheel about wheel axle (FLT1)
I wavelength of standing wave (L); half of projected contact
length (cf. fig. 9.5.1 of sec. 9.5)
M, (= My) rolling resistance moment (cf. fig. 8.1.1.) (FL)
M,.tJ moment acting from road to tire (cf. fig. 8.1.1), M, = align
ing torque (FL)
n frequency of motion (Hz)
pxj,j contact force per unit area acting upon tire in negative x, y
and z direction respectively (F/L1)
p, inflation pressure (F/L2)
qfj,j contact force per unit length acting upon tire in negative x, y
and z direction respectively (F/L)
r (or R) tire radius (L)
r, (or Rt) effective radius of rolling (L)
r, (or R,) loaded tire radius = wheel center height (L)
R radius of curvature (L)
s distance travelled (L); mode number
S tension force in tread band (F)
t time; pneumatic trail (= — M,/Ff) (L)
u tangential (longitudinal) deflection (L)
v lateral deflection (L)
i',,, lateral deflection of carcass and tread rubber respectively
(L)
V (= ds/dt) speed of travel (L/T)
VefJ creep (slip), rolling and sliding velocity respectively (L/T)
w radial deflection (positive outwards) (L)
W (-* - F,\ tire normal load (F)
X longitudinal horizontal force acting upon tire (F)
x,y, z coordinates with respect to moving system (fig. 9.5.1)
x, y, z coordinates of contact center C with respect to system fixed
in space
*« y*> 2, variation of wheel center position with respect to steady
state motion
*rt ya zc coordinates of contact point with respect to system fixed in
space
a slip angle (cf. fig. 8.1.1); crown angle
ft (= arctan dy/ds) path angle (cf. fig. 8.1.1)
546 MECHANICS OF PNEUMATIC TIRES

Y camber angle of wheel center plane (cf. fig. 8.1.1)


8 normal tire deflection (positive toward the center) (L)
A wavelength of motion (L)
ic longitudinal slip value (cf. eqs (9.4.27, 9.5.64)
KB.T percentage of brake and traction slip respectively
H coefficient of friction
(/> (= tty/ds) spin = yaw rate (RAD/L)
4- yaw angle (cf. fig. 8.1.1)
X deviation from steady state angle of rotation Qf (cf. fig.
9.4.13)
p mass density of tire tread band (FT2/L2)
a relaxation length for tire model without tread rubber (L)
a* relaxation length for tire model with tread rubber (L)
u frequency of motion (RAD/T)
u, reduced, spatial or path frequency (RAD/L)
SI speed of rotation of wheel (RAD/T)

8.2 Straight Line Rolling Experiments


H. C. A. van Eldik Thieme
In this section part 8.2.1 is devoted to a description of the equipment
used to measure forces and moments required for the investigations as
treated in sections 8.2 and 8.3.

8.2.1 Introduction to Total Force and Moment Measurements


Historical
A variety of measuring devices have been designed to determine the
forces and moments that arise in the tire contact area. Although initially
the interest was in load-deflection relationships in a radial direction, later
research emphasized the rolling resistance and the nonskid qualities of
tires (braking force coefficient). The need to evaluate the cornering force
coefficient and the slip angle was discussed as early as 1930 by Bradley
and Allan. The research in this area is well summarized in the master ref
erence list given in the papers by Milliken [1],1 Gough [2], and Kollmann
13].
Laboratory Tests
The majority of tests conducted to measure tire characteristics have
been run in laboratories, where the tire rolls on a large drum. However,
internal drum tire test machines are also available [3-5].
These indoor tests are useful for comparing the characteristics of differ
ent tires, but do not give exact information on the tire behavior on flat
roads because significant differences in the tire behavior are induced by
the curvature of the drum. Kollmann [3] and Krempel [4] have discussed
the effect of the radius of curvature of external and internal drums (fig.
8.2.1). They report that the advantage of an internal drum lies mainly in
the fact that it can be covered with any surface, including an exactly de
fined thickness of water.
1 Figure! in buckets indicate the literature references at the end of this section
MEASUREMENT OF TIRE PROPERTIES 547

fc R c

FIGURE 8.2. 1 . Schematic representation of test tires on external and internal drums.

To eliminate the effects caused by the curvature of the drum, a flat ro


tating disc has been used [3], but due to the finite width of the tire tread
the velocity of the inner and outer tire tread edge will be different, causing
slip in the contact area (fig. 8.2.2).
Another method of reproducing the traveling ground is the belt-type
tire tester, as shown in figure 8.2.3, consisting of an endless steel band or
conveyor belt running on two drums. The flexible belt is supported in the
tire contact area by an extremely stiff and thin air bearing, in order to
minimize the air gap variation due to the required variable vertical load
ing of the test tire. A drawback of such an air bearing is the large air con
sumption [6-8]. Calspan has available an advanced tire research facility
using appropriately surfaced steel belts simulating road surfaces from low-
to high skid resistance. Dry and wet pavement simulated conditions up to

testwheet

FIGURE 8.2.2. Test wheel on flat rotating disc.


548 MECHANICS OF PNEUMATIC TIRES

tettwheel

FIGURE 8.2.3. Endless belt tire tester.

full dynamic hydroplaning from low speed to 200 m.p.h. may be obtained.
A flat nozzle discharges a stream of water tangent to the belt as shown in
fig. 8.2.4.
Another indoor test is the method using a movable platform, as illus
trated in figure 8.2.5.
Because the movable platform is a low speed machine, it seems impor
tant to emphasize that as discussed not only surface curvature, but also
test speed produces an effect on tire force and moment properties.
The necessity to measure in a laboratory all forces and moments acting
on a tire, without excessive wear which would change the tire character
istics, dictated the design of various movable table machines [10-14].
These flat surface machines have the limitation of low speed, but afford
the easy control of test variables that can be only achieved in a laboratory.
Most tables provide for the possibility of bonding various surface materi
als, and others provide for a glass section for observing and photographing
tire deformations. Often the road surface can be wet, dry or iced at a con
trolled temperature for additional friction al studies [15]. Most machines

FIGURE 8.2.4. Belt tire tester equipped with a roadway watering system.
MEASUREMENT OF TIRE PROPERTIES 549

measure the forces and moments of a loaded, steered, cambered or


torqued tire. The results are often shown in the form of a carpet plot [ 1 2],
because this technique allows one variable to be plotted as a function of
two variables, so that accurate interpolation with respect to both inde
pendent variables is possible.
The Delft movable platform machine as shown in figure 8.2.5 is
equipped with an air spring device, in order to maintain a constant normal
tire load, as required for tests as described in section 8.3.2 (fig. 8.3.21).
The same machine can also maintain constant center height. The ma
chine is equipped with a slide mechanism and a turntable, both with grad
uated scales [14].
Laboratory testing machines and procedures for measuring the steady
state force and moment properties of passenger car tires have been pub
lished in S.A.E. Handbook supplement HS 210 (1975).

Towed Trailer Road Tests


The testing methods described above do not represent actual operating
conditions, and therefore devices were developed for measuring tire char
acteristics on road surfaces. The Air Force-Cornell tire tester [1] is well
known, consisting of a single-wheeled trailer towed behind a truck (fig.
8.2.6). The axle carrying the test tire is supported at its ends by units which
measure five forces, provision being made to include an instrumented
linkage for measuring the brake torque. The load cells are mounted on the
free ends of an U-shaped frame whose normal position is in a horizontal
plane. The mechanism permits the test tire to be steered and cambered 30
degrees and vertically loaded with a variable force up to 3,000 Ibs.

FIGURE 8.2.5. Movable platform machine with glass section to observe lire deformation.
550 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.6. Single wheeled trailer behind a truck for measuring tire characteristics.
A fifth wheel is also shown for accurate slip angle measurements.

Because the true slip angle applied to the tire consists of the steering
angle (tire relative to truck) minus the truck slip angle (truck relative to
direction of motion at test wheel), a fifth wheel unit was added to measure
the truck slip angle.
Another tire tester developed at the Delft Vehicle Research Laboratory
avoided the difficulty in measuring the slip angle by using two test wheels
in a trailer; the tire of the second wheel running at an equal but opposed
slip angle to the measuring wheel [10]. In this way the lateral forces in
duced by both wheels are now in equilibrium, resulting in zero slip angle
for the test trailer. Of course the motion of the trailer without test tires
must have zero slip angle, if accurately aligned and fitted with tires having
acceptable tire nonuniformity. A special air-sprung trailer (fig. 8.2.7.) with
a comfortable cabin for mounting auxiliary equipment was developed, in
cluding a six-component tire tester. In this trailer the tire load on the test
wheel is kept constant by means of an air spring, so that vertical move
ments of the trailer cannot influence the measuring results. The steering
system for varying the slip angle employs an extensive hydraulic regulat
ing apparatus, in order to rapidly adjust the various .slip angles desired un
der constant vertical tire loads. The camber of the tire tester can also be
adjusted. A water tank for spraying the road has been placed in the truck
towing the trailer.

Six-component Tire Testers


The paper in 1956 by Close and Muzzey [1] described one of the most
modern means of measuring tire characteristics, since it was the first com
plete machine to measure tire parameters on flat road surfaces at different
MEASUREMENT OF TIRE PROPERTIES 551

FIGURE 8.2.7. Towed trailer road tests with a six-component tire tester. The test wheel is
mounted in the airsprung trailer.

FIGURE 8.2.8. Forces and moments of a cambered wheel with enlarged view of the contact
area.
552 MECHANICS OF PNEUMATIC TIRES

speeds and loads. Most available tire testing equipment was capable of
measuring the steady state characteristics of the tire, but for steering sys
tem stability problems the time or distance behavior of the tire must be
known.
The obvious advantage of a six-component tire tester is not only its abil
ity to measure tire characteristics on flat roads, but also its utility as a pre
cision laboratory test machine on a drum.
In this way it is possible to compare test results obtained from road and
drum tests with the same tire.
The interpretation of forces acting in the tire road contact area has been
discussed by Fonda, Close and Muzzey [1].
In the contact area of a cambered tire there is a distribution of forces
normal to the road plane and a distribution of shearing forces in the road
plane. These forces can be resolved into three forces and moments acting
on the tire at the so called center of tire contact C.
Figure 8.2.8 shows a cambered wheel in contact with the ground, and an
enlarged view shows the forces at a point x, y in the area A.
The equations are given by:

Fx M,-tydF,
'A

F, = I dFr M, = I - x • dF,
JAA JA

F,-t dF, M,= f (x-dF,-ydFJ.


•A 'A

The vector of the pressure in the positive direction exerted by the tire
upon the road is denoted px, p^ p,. Thus px is the force per unit area, and
p., • dA = — dFx, py · dA = — dFy, pz · dA = — dFz and therefore

The choice of the location of the origin Cw of an additional set of axis


x*y*z* through the center of the wheel reduces misinterpretation. Figure
8.2.9 shows the new axis, and after transformation we obtain
F* = Ft M? = M,-FyRcosy- F,Rsin y
F,*-F, M* = M, + F,R cos y
F.* = F, M* = M, + F,R sin y.
Because the load cells make their measurements with respect to the wheel
axis, a third set of axis is shown in figure 8.2.10, and, after transformation,
the equations read:
F* = ¥„ M* - MIW
F* = F^ cos y - Ffw sin y M* = M^ cos y - M,w sin y
F? -= F,w cos y + Fyw sin y M* = M,w cos y + M^ sin y.
MEASUREMENT OF TIRE PROPERTIES SS3

FIGURE 8.2.9. Transformation offorces and momentsfrom contact center C offigure 8.2.9 to
center of the wheel C,,

FIGURE 8.2.10. The load cells make their measurements with respect to the wheel axis
requiring the transformation offerees and moments to a third set of axes (X^, Y*. ZJ.
554 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.11. The possibility of mounting the tire at a free end of the measuring axle is
shown.
The camber of the lire leuer can be adjusted.

To provide for easier changing of the tire to be tested than in the Cornell
machine, a somewhat different design was chosen for the Delft tire tester
built in 1959, [14], as shown in figure 8.2.1 1. The advantage of this design
is that of mounting the tire at the free end of the measuring axle.
Figure 8.2.12 illustrates the relationship of wheel axis forces and mo
ments to measuring devices P, fl, R and a flexible coupling S. The output
readings can be used to compute the externally applied forces and mo
ments for the coordinate axis shown in figure 8.2.12.
F,» = /„ + /„ A/,.. = bf;r - a/,,

To calibrate the tire tester a special very stiff test rig was built (fig. 8.2.13)
having three absolutely perpendicular axes, the tire tester being accurately
fixed to an extremely stiff bed plate. With the aid of three dead-weight
controlled dynamometers, several combinations of forces and moments
can be induced in the tire tester, giving the required output corrections.
With the test rig it is also possible to prove that, for instance, a static force
in the x-direction will not influence the output of strain gage bridges in
other directions. Not only is static calibration required, but also assess
ment of the dynamic behavior of the measuring unit is necessary, in order
to define its limitations.
Considering the automobile as a two degree of freedom system, we ob-
MEASUREMENT OF TIRE PROPERTIES 555

FIGURE 8.2.12. Relationship of wheel axis forces and moments for the Delft tire tester.

FIGURE 8.2.13. Test rig having three perpendicular axis to calibrate the Delft tire tester.
556 MECHANICS OF PNEUMATIC TIRES

serve a low body frequency (0.8-2 Hz) and the high axle frequency (10-16
Hz).
Influence of any of these vibrations on actual force variations of the tire
requires a tire tester with reasonably high limiting frequencies. After
mounting the tire tester in a rigid structure, excitation was impressed with
an excitor in all three mutually perpendicular directions. Experiments
were made with a loaded tire as well as with wheel and tire removed, the
first resonant peak being observed 55 Hz, and flat response curves were
obtained up to 30 Hz, deviations beginning at 35 Hz. An approximate
method was employed to calculate the resonant frequencies of the com
plex system, and yielded a frequency spectrum close to that determined
experimentally. Filters with suitable cut off frequencies were available to
eliminate disturbing resonant frequencies [17].
Mobile high speed tire tester.
At the Vehicle Research Laboratory of the Delft University of Tech
nology in the Netherlands an advanced test vehicle for high speed tire
traction research and dynamic response of tires has been built, in order to
obtain vehicle oriented tire data.
Since transient tire properties are highly affected by speed and rate of
change of slip angle, it is important that tests be conducted at realistic
highway speeds and on real road surfaces. This new mobile tire tester will
be a powerful tool for the investigation of the dynamic behaviour of car
tires on different road surfaces at speeds exceeding 100 km/h. It is possible
to measure transient tire data for simulation, with rates well within the
range of most passenger car maneuvers and related to the actual steering
process. The dynamic effects on a simulated test course can be measured
in order to obtain slip angle and load-time histories in various maneuvers,
e.g. by simultaneous variation of slip angle, vertical load and braking
force; these variations are not uncommon in accident avoidance maneu
vers.
The test vehicle is a front wheel drive passenger car which has been
modified so as to house a measuring system. The prime feature of this sys
tem is a compact and light weight measuring hub using piezoelectric force
transducers which have a very wide measuring range up to Fz = 10 KN. In
comparison to conventional measuring hubs, the light weight construction
of hub and suspension system leads to a lowest resonant frequency of 10
Hz, which enables a better simulation of dynamic wheel load variations of
passenger cars. The lowest resonant frequency of the measuring system is
of the order of 1 k Hz, thus enabling accurate measurements of the tire
force transients. The lay out of the measuring system in the test vehicle
may be seen in figures 8.2.14 a/b. In tests for measuring the lateral slip
characteristics, the side force of a dummy wheel compensates for the side
force developed by the measuring wheel. The lateral stability of the test
vehicle is thereby preserved. The measuring hub is shown in Fig. 8.2.15.
The test wheel can be steered and braked with electro-hydraulic servo
valves, as shown in figure 8.2.16 while all the forces and moments are con
tinuously registered with the help of a digital casette recorder using pulse
code modulation technique. The maximum vertical load on the test wheel
is about 3.5 kN when the dummy wheel is also used, but in special low
speed testing a load of 7 kN can be applied. The usual tire sizes in the
12"- 15" rim diameter range can be measured. The maximum test speed is
150 km/h. A watering system is available for measurements on wet roads,
and in order to measure tire hydroplaning characteristics.
MEASUREMENT OF TIRE PROPERTIES 557

,
I
I

w
a:
0
558 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2. 14b. Mobile high speed tire tester

Road Platform Tire Tester


An apparatus for measuring the total forces exerted on a road surface
by the wheels of a moving vehicle can be of considerable interest, because
it gives information about load transfer, cornering and traction or braking
forces of a vehicle passing over the measuring platform [14].
A photograph of such a platform with a steel surface table is shown in
figure 8.2.17. There is the possibility of bonding various surface materials
to the table. The platform is supported by strain gaged bars operating in
bending, the strain gages being bonded to the bars in such a way that only
shear and vertical forces are measured [18-19].
An approach such as this is satisfactory when measuring total tire forces
at low car speeds, because the output signal obtained is not strongly influ
enced by the excitation of the lowest natural frequency of the platform at
about 50 Hz. Figure 8.2. 1 8 shows in dashed lines the supposed trapezoidal
load of a tire moving onto, over and off the platform, as well as the output
signal obtained at speeds of 25 km/hr. is also illustrated; it is seen that it is
impossible to find the trapezoid from the output signal obtained. How
ever, using a filter with an inverse transfer function of the measuring plat
form [20-21], satisfactory results could be obtained as shown in figure
8.2.19. This example may illustrate that a platform originally intended to
calibrate a mobile tire tester on the proving ground, with certain limita
tions, could also be used as a tire tester.
A better approach is, of course, to raise the stiffness of the measuring
system by replacing the bending elements supporting the table by bars in
MEASUREMENT OF TIRE PROPERTIES 559

compression or tension [22], or using piezoelectric force transducers. The


complicated inverse filter measuring system may be omitted.
Besides the possibility of measuring three perpendicular forces Fx, Fy, Fz
of a tire moving over a road platform, the position and path of the test
wheel with respect to the platform axes have to be known, as will be
treated in the sections 8.3.3 and 8.3.5.
8.2.2 Load-Deflection Relationships
In this section the vertical load deflection characteristics will be treated.
In addition, some attention will be paid to the longitudinal (tangential) be
havior, and to a lesser extent the lateral and torsional characteristics of the
tire will be discussed. The latter two types of deformation are directly re
lated to cornering and yaw response of the tire treated in sections 8.3 and
9.5.

FIGURE 8.2.15. Delft light weight measuring hub using piezo-electrtc force transducers.
560 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2. 16 The test wheel can be steered and braked with electro-hydraulic servo valves.

Vertical Load-Deflection Relationships


The lowest natural frequency of the vertical vibrations of an automobile
is determined mainly by the sprung mass and the stiffness of the wheel
suspension springs. This frequency normally ranges from about 1 to 3 Hz.
The second resonant frequency is determined by the unsprung mass and
the radial tire stiffness. It usually lies in between 8 and 20 Hz. The radial
MEASUREMENT OF TIRE PROPERTIES 561

FIGURE 8.2.17. Road platform tire tester having the possibility of measuring three
perpendicular forces of a tire moving over the platform.

kgf
2Skm/h
wheel load F,
output signol

0 30 70 100 millisec.

FN
100km/h

mi III tec.
7,5 K

FIGURE 8.2.18. Response curves of the road platform tire tester for vertical tire load at two
speeds.
562 MECHANICS OF PNEUMATIC TIRES

Ml

HI
FIGURE 8.2. 19. Illustration of results obtained with the road platform tire tester by adding a
filter with an inverse transfer function.

tire flexibility influences the vehicle vertical motion and the time variation
of the vertical tire load. The former aspect is important for ride quality,
whereas the second aspect indirectly influences the directional stability of
the vehicle. Besides, tire flexibility plays an important role in the life of ve
hicle components and road surface.
When a standing tire (nonrolling, V = 0) is loaded vertically and after
wards unloaded, a loop in the load-deflection curve is observed, which is
due to hysteresis and friction losses [14]. Investigation with a tire rolling
on the inner surface of a rotating drum of 3.8 m. diameter [5], shows that
the loop area decreases with increasing rolling speed as in figure 8.2.20. It
has been generally accepted that the damping produced by the rolling tire
is very small, and can be neglected in studies dealing with vertical axle
motions of relatively low frequencies. Results obtained with the Delft tire
tester are in conformity with the above statement [10]. The curves of figure
8.2.20 furthermore indicate that for the bias ply tire under consideration
an increase of the vertical stiffness occurs with increasing speed. The be
havior of a radial ply tire shows a similar trend and has been illustrated in
figure 8.2.21. The shape of the curves, however, are different [5]. Typical
for bias ply tires is the double curved shape. The radial ply belted tire usu
ally shows a progressive load-deflection curve as represented in figures
8.2.22-23. These results were obtained with the Delft tire tester on a drum
of 2.5 m. diameter [23-24].
The inflation pressure />, has a marked effect on the radial stiffness. For
a nonroUing bias ply tire the curves shown in figure 8.2.24 have been ob
tained [5]. Similar results are shown for a 165-13 tire rolling at 5 km/hr on
a steel drum of 2.5 m. diameter [24].
The tire stiffness is composed of a more or less constant part originating
from the structural rigidity of the tire, and of a part dependent on the air
pressure pi [25-26]. Under rated conditions, the latter part is predominant.
It has been found that the structural rigidity of the aircraft tires may sup
ply 3 to 8 percent of the total stiffness, whereas for automobile tires this
percentage may amount to about 15 percent [26-28]. These relatively low
percentages explain the considerable influence which the inflation pres
sure has upon the radial tire stiffness, as indicated in figure 8.2.24a.
According to measurements of Henker [5] the radial stiffness of the tire
under consideration becomes approximately 20 percent lower when the
tire rolls upon an external drum surface, as compared with an internal
drum surface of the same diameter. This means that when the tire rolls
MEASUREMENT OF TIRE PROPERTIES 563

V=0 km/'h
1- V=50km/h
V=100km/h

Tin dtrttction z,

FIGURE 8.2.20. Vertical load deflection relationships at various speeds for a bias-ply tire,
obtained on an internal drum machine.

ViOkm/h
VsSOkm/h

10 20 30 50 mm

Tirt dtfltction Z,
FIGURE 8.2.21. Vertical load deflection relationships at two speeds for a radial ply tire
obtained on an internal drum machine.
564 MECHANICS OF PNEUMATIC TIRES

bias ply 175 S.U


Pisl5kgf/cm«

20 30 10 mm
tire deflection z.

n«il ply 175 SRU


600 * Pi: 1.9 kgf/cm>

500

uf

20 30 40mm
tire deflection za
FIGURE 8.2.22. Vertical load deflection relationships al various speeds obtained on an
external drum machine for a bias ply and radial tire.
MEASUREMENT OF TIRE PROPERTIES 963

Q, Mot ply
V km/h
rodlolpty

20 30 40 so to 70 mm
tin difltctien i.

FIGURE 8.2.23. Demonstration of the difference in behavior ofa radial ply and bias ply lire at
various speeds.

over a flat surface, the stiffness is expected to become approximately 10


percent lower than values obtainable from the internal drum machine.
It has been often observed that characteristics for both radial and bias
ply tires can be approximated very well by straight lines passing through
the origin. Only for relatively low values of vertical load does a noticeable
deviation from linearity take place. The slope of these straight lines deter
mines the vertical stiffness of the tire, which consequently is assumed to be
independent of load in the practical range of interest.
From tire technical data books the unloaded radius as well as the
loaded radius can be gotten as can the maximum load at several inflation
pressures. From these data a vertical spring rate C1 may be approximated.
The measured real values C2 are generally about 5-20% higher for pas
senger car tires and 10 to 40% for truck tires. This is due to the relatively
high deflection at loads from zero to small values, as illustrated in figure
8.2.24c.
Figure 8.2.25 presents the results obtained for the spring stiffness of the
tire as a function of speed for a number of values of inflation pressure [29].
It may be noted that according to these experiments, the stiffness of the
bias ply tire is very sensitive to the speed of travel. Similar results were ob
tained with the Delft tire tester as shown in figure 8.2.26. Due to the in
crease in observed radial stiffness an increase in loaded tire radius (wheel
center height) will occur. This increase with speed will become even larger
due to the phenomenon of tire growth. Figure 8.2.27 shows the increase of
the free tire radius R due to tire growth, and the increase of the loaded
radius Rt due to both tire growth and stiffening of the tire [29].
An alternative method for the assessment of the radial stiffness consists
in determining the resonant frequency of a single mass spring system with
known inertia and spring stiffness of applied load, in series with the resil
ient tire rolling on a drum or belt [30]. A schematic of this drum type ma
chine has been shown in figure 8.2.28 operating in the frequency range of
10-15 Hz. The value of the spring rate obtained by this method is called
the dynamic tire stiffness, and it is shown that the dynamic stiffness de
creases sharply as soon as the tire is rolling. Beyond a speed of about 20
km/hr. the influence of speed becomes less important, as demonstrated in
figure 8.2.28. These findings are in contrast to the results as shown in fig-
566 MECHANICS OF PNEUMATIC TIRES

W 10 JO to mm
Tirt def lection Z0

FIGURE 8.2.24a.

10 20 30 40 SO CO 70mm

Ttrt deflection la

FIGURE 8.2.24b.
Vtrtical spring ritt

Tire dtfltction 2t

FIGURE 8.2.24c.

FIGURE 8.2.24. Variation of vertical load deflection relationships as a function of the


inflation pressure pf
MEASUREMENT OF TIRE PROPERTIES 567

kff/cm*

iZpkjf/cm1

SO 100 ISO 50 WO ISO


Ipttd of travel (km/h) < ipttd of travel (km/h)

FIGURE 8.2.25 The vertical stiffness or spring rate as a Junction of speed at various inflation
pressures for a bias ply and radial ply tire.

ure 8.2.25 and figure 8.2.26, obtained from the slopes of the load deflec
tion curves at different speeds of rolling. For amplitudes within the range
investigated, up to 10 mm., the stiffness of the rolling tire remains nearly
constant, whereas a nonrolling tire shows a nonlinear decrease of the stiff
ness with amplitude [3 la].
The spring rate also plays an important role in the vibration transmis
sion properties of tires. The vertical transmissibilities of different tires can

x bios ply 175S.U


a radial ply 175 SRU
.335 kgf

20 40 60 80 100 120 UO km/h


sptcd of travel

FIGURE 8.2.26 The spring rate as a function of speed, demonstrating the difference in
behavior of a bias ply and a radial ply tire.
568 MECHANICS OF PNEUMATIC TIRES

FN • 300 kgf
LS kgf/cm»

150 km/h so WO 150 Km/h

FIGURE 8.2.27. TVie increase of the free tire radius at FN = 0 kgf as a function of speed, and
the increase of the loaded radius at FN - 300 kgf as a function of speed for a bias ply and a
radial ply tire.

be compared by using a shaker platform [31b] and denning the trans-


missibility as:
axle force amplitude
transmissibility =
foot print displacement amplitude

Figure 8.2.28 shows the transmissibility (spring rate) of a rayon bias tire, a
polyester glass belted bias tire, and a steel belted radial tire, all in the 1178-
1 5 size. These well-known very different transmission properties resulting
from structural differences of bias and radial tires will be treated in detail
in section 9.4.2. For further information on tire spring rate the reader is
referred to the literature [31c].
The influence of preload, inflation pressure, rim width, and cord angle
has been discussed in reference [3 la], the influence of, among other things,
speed and drum curvature in references [5.32] and of tire size in references
[33-34]. In section 9.4.1 of this book the tire loading process and the envel
oping properties of the tire have been analyzed.
Horizontal Load-Deflection Relationships
The literature on the horizontal elastic characteristics of the nonrolling
tire is scarce. Some of our own unpublished investigations will be dis
cussed below.
The same tires as in figure 8.2.23 and figure 8.2.26, have been examined
under rated conditions, and are of the radial and of the conventional bias
type.
The tire is now loaded upon a flat cast iron plate, in contrast to the tests
described previously. The normal load has been kept constant during each
test. The measured vertical stiffness of the nonrolling tire amounts to ap
proximately Cz = 17 and 20 kgf/mm. respectively. Except for the longitu
dinal stiffness characteristics, all force and moment measurements have
been executed with the six-component measuring hub.
The longitudinal stiffness Cx of a tire is important for the study of longi
tudinal isolation of disturbances caused by road irregularities. The longi
tudinal stiffness may be determined in two ways which do not necessarily
give the same results. With the first method, the wheel is fixed in space and
the surface upon which the tire is loaded is displaced in the longitudinal
MEASUREMENT OF TIRE PROPERTIES 569

drum diom»t»r 2,5 m


amplitude 5-10 mm.
tlr»s 185 » 400

CO
C
>%
«

10

0 5 10 20 30 40 50 60 70 100 km/h
speed of travel

FIGURE 8.2.28a. TTie decrease of the vertical dynamic spring rate as a function of speed.

direction as shown in figure 8.2.29. With the second method, the surface
and the wheel axle remain fixed, while the wheel is rotated by a torque
about the wheel axis. The resulting tire deformations are different in these
two cases. The results obtained with the second method are presented in
figure 8.2.30. The deformation rate of all tests was low. The rotation of the
wheel has been increased until complete sliding takes place. The longitu
dinal (tangential) stiffness of the radial ply belted tire appears to be much
lower than the stiffness of the bias ply tire. At zero longitudinal dis
placement of the wheel at road level, we obtain for the stiffness approxi
mately C. = 27 and 45 kgf/mm. respectively [35].
The lateral stiffness Cy of a nonrolling tire is important for the relaxa
tion properties of the rolling tire. In combination with the cornering stiff
ness CFα, to be discussed later, the so-called relaxation length can be deter-
570 MECHANICS OF PNEUMATIC TIRES

BIAS
— BELTED BIAS
— • RADIAL

20 40 60 80 100 120 140


FREQUENCY (Hz)
FIGURE 8.2.28b. Transmissibility (spring rale) variations over frequency of input for bias,
belted bias, and radial tires, all in H78-15 size.

mined (eq 9.5.51)). The lateral compliance may furthermore give rise to a
lateral and yaw vibration of the vehicle with respect to the lower tread
portions of the tires.
For the same radial and bias ply tires figure 8.2.31 shows the measured
lateral force-displacement characteristics. The flat cast iron plate on which
the tire is loaded has been moved sideways, until complete sliding takes

FIGURE 8.2.29. Measurement of the longitudinal stiffness.


The left wheel Axed in spice is loaded with an identical lire The force required lo displace the sleelplate u measured with
a dynamometer.
MEASUREMENT OF TIRE PROPERTIES 571

kgf

V x/
W97 /////

_ P.-J3JJS^/cmt ^.
s'
B 7 /< ^ lf/cm«

bias ply 175 51


9
^ // / radial ply 175 SRU -
/ F.. 335 kgf
° 100

I
•ngular lisplacement ) . do'2 •ad)
/' 3 6
2
9
:I
12
;
15
s
longitudinal displacement of wheel
18
i '
21 mm
1

at road level RLX (mm)

FIGURE 8.2.30. The tangential stiffness of the radial ply tire is much lower than the bias ply
' tire.

place. Again the radial ply tire shows a lower stiffness than the bias ply
tire. We obtain at zero displacement approximately Cy = 9 and 12 kgf/
mm. respectively.
Similar results were obtained by another method as reported by the au
thor [10].
The lateral stiffness is generally in the order of 50% of the vertical stiff
ness.
The torsional stiffness Q of a nonrolling tire about its vertical axis can
be used in combination with the self-aligning torque rate CMα for the de
termination of the longitudinal tread stiffness parameter κ* (eq (9.5.120)).
The torsional stiffness may furthermore be of value for the assessment of
the steering wheel torque required at zero or nearly zero forward speed
(parking).
The characteristics obtained are presented in figure 8.2.32. The turning
of the wheel about the vertical axis has been continued until full sliding
with respect to the flat cast iron plate occurs. Similar to the tests discussed
above, the radial tire appears to be more compliant also in this respect. We
find at zero angle of rotation approximately Q = 250 and 290 kgfm/rad
respectively [35].
In the past similar tests have been executed on the movable platform
machine, as illustrated in figure 8.2.5 of section 8.2.1.
572 MECHANICS OF PNEUMATIC TIRES

500

400

Pi.ljskgf/cm'
C 300

_ 200

bias ply 175 514


100 radial ply 175 SRK
F=335kgf

10 20 30 40 50 60 70 80 mm
lateral displacement

FIGURE 8.2.3 1 . The lateral stiffness ofthe radial ply tire is much lower than the bias ply tire.

20 bias ply 175 S14


radial ply 175 SRU
F,335kgf

Pi.1.5kgf/cm«

10

02 (rad)
_PJ_
3 5° 10e IS"
angle of torsion 41

FIGURE 8.2.32. Comparison of bias ply and radial ply tires in their torisonal stiffness
behavior.
MEASUREMENT OF TIRE PROPERTIES 573

Actual Vertical Dynamic Tire Forces


The measurement of dynamic tire force has been required for various
purposes, such as car handling, optimizing shock absorber settings, studies
of what is decisive in respect to dynamic stress imposed upon the wheel or
road when passing over wavy short unevennesses, analysis of enveloping
forces of tires passing obstacles, etc.
The dynamic tire force F, consists of a constant part Fz0 and a variable
part Ft. Consequently, we have:
F= F + F
*• M * to ^ •* *•

Considering a very simple two mass system such as shown in figure


8.2.33, it seems very easy to find the tire force variations F, from:

The measurement of the accelerations of only the car body and the axle
are required. But with such a simplification it is very difficult to calculate
the reduced masses involved accurately enough [36-37]. Even with a sup
posedly rigid car body the actual vibrations are much more complicated,
due to rolling, pitching, bouncing, etc. Acceleration measurements usually
fail to give results accurate enough to yield useful data on actual tire
forces.
To measure the dynamic tire force Fz during road tests, the tire is con
sidered to have no damping. We obtain t. by measuring the tire deflection
through the variable distance between axle center line and the ground,
and by assuming a known value of the radial stiffness Cz. It follows from
figure 8.2.33 that:
F, = C,(zc - za).
The radial tire stiffness Cz may be obtained as a function of speed from
laboratory drum tests. This test method can be refined by using an elec-
trohydraulic vibrator to shake the rolling tire at different amplitudes and
frequencies [23-31].
In other cases the frequency sensitivity is obtained by mounting the test

""b

FIGURE 8.2.33. Model of two mass spring system passing over an uneven road surface.
574 MECHANICS OF PNEUMATIC TIRES

wheel in a stiff frame with an actuator plate moving the footprint up and
down. But it should be noted that results in this case are obtained with a
non roll ing tire, and the influence of the deformation of the cross section of
a rolling tire due to centrifugal forces cannot be taken into account. It will
be very difficult to correct for an "effective" tire mass in this case. Having
discussed the radial tire stiffness, we observe several methods of measuring
the tire deflection.
The Miihlfeld antenna system [38] measures the variable capacity of a
moving condensor plate above the ground, and requires a plate area of
about 25 X 25 cm. As a consequence, dynamic tire loads due to road un-
evenness of half wavelength less than 25 cm. cannot be accurately mea
sured. There is thus some uncertainty of the value of the dynamic tire stiff
ness when passing over a short obstacle.
Another system introduced by von Bombard [39, 90] measures the vari
ation of the width of a cross section of the tire. Later modifications having
sensing devices with a roller assembly measure the position of a point on
the left and right hand side of the tire wall as in figure 8.2.34. The buffing
ribs of the side walls in contact with the rollers have to be ground to cor
rect for the lateral tire nonuniformity. The normal procedure of averaging
left and right hand displacements of the tire wall as a mesaure of the verti
cal load gives erroneous results when rolling under a slip angle. It is seen
from figure 8.2.35 that the lateral and radial displacements of a point on
the tire wall are very much dependent on the value of the slip angle [40].
The method of making these measurements is illustrated in figure 8.2.36.
As can be seen from the photograph, a point on the tire wall is contacted
by two spring loaded wires, which are perpendicular to each other, the
wires being attached to potentiometers. This method, shown in figure
8.2.34, can only be used in straight line rolling on relatively smooth roads
when no lateral forces are present, and corrections have to be made for the
speed dependent character of the tire width.
The radial and lateral displacement of a point on the tire wall is also
very much dependent on the position in degrees from the contact center
(fig. 8.2.37) due to the buckling effect in the center of the contact zone [40].
Attempts have been made to measure the tire deflection at a point in the
wheel center plane at the inner liner of the tread when passing through the
contact zone (fig. 8.2.38). A spring loaded flexible cable was bonded to a
piece of canvass attached to the inner liner of the tire, while the other end
of the cable was connected to a potentiometer mounted in the wheel rim
of a tubeless tire. The restraining spring was relatively stiff in order to give
a frequency response compatible with the large accelerations of the tread
when passing through the center of the contact zone. Because the poten
tiometer reading had to be corrected for the speed dependent character of
the tire tread deflection, and on rough roads the slider vibrated upon the
potentiometer windings, this method appeared to be too complicated and
unreliable.
Various other sensing devices have been proposed to measure the tire
deflection of a point in the wheel center plane at the inner liner of the
tread when passing through the contact zone. Such devices are based on
the measurement of the variation of the resistance [40-42], capacitance or
inductance, on photocell counter devices for light beam pulses, on the de
tection of the variation of the radioactivity of an emitter plate mounted in
the tread, etc., but all methods apparently are without success under con
ditions of a combination of traction and cornering [40].
Another method used the measurement of the variation of the signal
MEASUREMENT OF TIRE PROPERTIES 575

FIGURE 8.2.34. Method of obtaining the vertiral dynamic lire force by measuring the
variation of the width of a cross section of the tire.

•T +T 0* \-3* .6' slipangU

FIGURE 8.2.35. Influence of the slip angle on the lateral and radial displacements of a point
on the tire wall.
576 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.36. Method of measuring the lateral and radial displacements of a point on the
tire wall

Radial contact center


Displacement

rotation angle
* degrees

Lateral 7.50-U
Displacement
Load 300 kgf
2ipsi
FIGURE 8.2.37. Variation of the lateral and radial displacements ofa point on the tire wall as
a function of the wheel rotation through the contact center.
MEASUREMENT OF TIRE PROPERTIES 577

FIGURE 8.2.38. Measurement of the radial tire deflection with a spring loaded cable
connected to a potentiometer.

from a strain-gaged symmetrical rim, but due to the momentary position


of the strain gage bridge the load variations cannot be accurately mea
sured when passing over an obstacle [43-44].
The method of measuring inflation pressure changes has also been re
ported for a tire rolling over a cleat on a test wheel [45]. The very small
change in inflation pressure of the tire was detected with a differential ma
nometer mounted outside the test tire, and this pressure change can be re
lated to the process of envelopment of the obstacle. A similar method was
adopted with a sensing device mounted at the rim inside a tubeless tire.

8.2.3 Effective Rolling Radius


Introduction
An analysis of the deformations of a rolling tire shows that the effective
roiling radius Re is the ratio of the linear velocity V of the wheel center in
the X-direction to the angular velocity il of the wheel:

The effective rolling radius is often erroneously assumed to be equal to the


center height h or so called loaded radius Rh being the distance from the
center of tire contact to the wheel center. When R is the undeflectcd free
tire radius, it will be shown that:

that is, the rolling tire travels farther per revolution than determined by
578 MECHANICS OF PNEUMATIC TIRES

using its center height as rolling radius, but less than described by the free
tire radius.
A rolling wheel may be considered as a wheel rolling along the station
ary ground, or the center of the wheel may be taken as fixed, so that the
ground moves relative to it. Both cases will lead to correct results and the
hitter case is represented in the flat platform machine.
Consider figure 8.2.39, which represents a tire on a flat platform ma
chine with a number of equally spaced radii, drawn on the surface of the
undeflected free tire, dividing the circumference into n equal elements,
each with length l = -.
n
It is observed from the photograph that tread elements in the zone im
mediately before contact are compressed, as well as elements in the area
just after contact has been lost (fig. 8.2.40). Due to compression the tread
elements in the contact zone are shorter, and therefore in - of a revolution
n
of the wheel, the platform moves a distance λl, which is less than length

l = -. This is the same as saying that when the platform moves a dis-
n
lance 2irR, the wheel will turn more than one revolution.
From tests with a nominal loaded tire under straight line rolling condi
tions we also observe very little longitudinal slip in the contact area (re
sulting in almost no wear), so that the compressed tread elements will
travel with the platform speed V.
When the circumferential speed of the undeflected upper part of the tire
is called V0, we find [24]

"}
R.
This was demonstrated experimentally by several authors [46-47] and is in
accordance with the situation shown in figure 8.2.41, where the wheel cen
ter 0 is maintained fixed and the platform, in position 1, moves to the right
over the distance X. Then the point of the tire surface initially at C1 moves
to £),. Simultaneously, the point of the tire at the rim of the wheel A moves
to B. The wheel thus rotates over the angle AOB = <p. The distance be
tween the center of the wheel and the platform in the position 0 (no load)
and position 1 (loaded) is OC0 = R and OC1 = R1 respectively [47]
C,D, = X=R,-tp
C,£, = X, = R tan <p
A'o = R sin <p
R. = —
whence — X --Tand
tan op . —
R. = —
X --*•
sin o>
K, A, <p A AO qp
MEASUREMENT OF TIRE PROPERTIES 579

FIGURE 8.2.39. Measurement offree radius R, loaded radius R, and effective radius R, on the
movable platform machine.

h.R,
Compression

1
1 V. Platform
[Contact Icnglh
1
1
1
1

^
1
V.flh
Ti
ii
Tirt circumference

FIGURE 8.2.40. Circumferential speed of a tread element when passing through the contact
zone.
580 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.4 1 a. Deformation ofa tire on a movable platform when a radial load is applied.

center
of rotation
FIGURE 8.2.4 lb Due to the highly non-linear character of the normal force-deflection curve
in the region of loadsfrom zero to small values, the rigidity ofside walls and tread band play an
important role. The tread thickness d will influence the deformation process in the contact area
in the neighbourhood of zero load to a large extent, and it may even be observed that in this
region of very small loads if' becomes negative, that is R, < R/ This may be explained by in
troducing an effective road level at a sufficient distance above the actual road surface. The effec
tive road level will be located above the center ofrotation in order to have a positive longitudinal
slip velocity on this level.
MEASUREMENT OF TIRE PROPERTIES 581

For <p not too large, tan <p/<p is slightly larger and sin <p/<p is slightly
smaller than unity.
Instead of an effective rolling radius, an effective deflection δeff = R - Re
can be introduced, which differs from the actual deflection δ = R - R,.
The effective rolling radius can also be expressed as

where TJ' indicates position of the center of rotation P [29] (fig. 8.2.4 lb).
See also section 9.4.1, Fig. 9.4.10.
Due to the highly non-linear character of the normal force-deflection
curve in the region of loads from zero to small values, the rigidity of side
walls and tread band play an important role. The tread thickness d will in
fluence the deformation process in the contact area in the neighbourhood
of zero load to a large extent, and it may even be observed that hi this re
gion of very small loads η′ becomes negative, that is Rt< R1. This may be
explained by introducing an effective road level at a sufficient distance
above the actual road surface. The effective road level will be located
above the center of rotation in order to have a positive longitudinal
slip velocity on this level. Fig. 8.2.4 Ib. For zero load it may be assumed R
= R.= R1.
Methods of Measuring Effective Radius
A. Movable platform (fig. 8.2.39)
Flat surface machines have the limitation of low speed, but good control
of variables can be achieved.
The radial load is applied and measured by means of a load cell, the
free radius R and loaded radius Rl are measured with a scale.
When the platform is moved over a measurable distance X, the corre
sponding angle of rotation <p of the wheel is determined by means of a
scale and pointer or other suitable instrument attached to the supporting
structure and wheel rim respectively.
The effective rolling radius is:

/?,= 180 —
mp

when </ is measured in degrees.


B. Drum tests
After conditioning the test tire at the required load, inflation pressure
and speed, the effective radius is determined by the ratio of the circum
ferential drum speed V and the angular velocity Ω of the test tire,
R - V •
R'~~
The angular velocity is accurately measured with a photocell counting
device as in figure 8.2.42, which may be equipped with continuous record
ing of the angular velocity [10, 24].
The variation in the effective rolling radius per wheel revolution due to
tire nonuniformities can also be measured with drum tests (sec. 8.2.6).
Nonuniformities produce angular wheel accelerations that result in tan
MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.42. Photocell counting device for measurement of angular velocity.

gential force variations as measured with a tire tester or uniformity ma


chine.
The free radius R of the tire can be measured at a given speed through
skimming contact with the drum, by measuring the distance between
wheel center and the surface of the rotating drum. (See also tire unifor
mity machine, sec. 8.2.6 (fig. 8.2.81)).
C. Road tests
Road tests have been reported [48] by driving an automobile at walking
speed over a measured distance and counting the number of revolutions of
the free rolling front wheel. Tests were run with several tires at various in
flation pressures.
Parameters Affecting the Effective Radius and Loaded Radius
A. Carcass construction
Factors influencing the deformations of the rolling tire, such as carcass
construction, tread thickness, number of plies, cord type and cord angle,
rubber compound, etc., will also influence the effective radius.
Experimental results, obtained with the Delft tire tester mounted in a
rig with the loading device above a drum 2.5 m. in diameter, show the dif
ference in behavior of a conventional bias compared with a radial-belted
tire.
Using Re = XR, the conventional bias tire shows a value of \ = 0.95 and
h = R1 = 0.93 R for nominal load.
For the radial ply construction with an inextensible belt, having less
compression of the tread elements when passing through the contact zone,
values of λ = 0.96 were obtained. The radial deflection of this tire is more
than in the former case, giving a value of R1 = 0.91 [24].
MEASUREMENT OF TIRE PROPERTIES 583

Bias ply Spied


175 S.U 120 km/h
P| = 1,8 bar — — —— 60 km/h

650 1550 2450 3350 N


Tin load Fz

FIGURE 8.2.43a Effective and loaded radius for a 175 S-J4 bias tire.

B. Effect of tire load and inflation pressure


An increase in load at constant pressure results in a larger deflection of
the tire. Due to an increase of the compression of the tread, a decrease of
the effective radius is observed shown in figures 8.2.43a and 8.2.43b.
A decrease of inflation pressure has the same effect as an increase of tire
load.
It is observed that for both radial and bias ply tires the effective radius
decreases with load much less rapidly than the loaded radius, and both
variations are nearly linear.
The influence of the tread thickness for a steel radial tire is clearly
shown in figure 8.2.44.
An interesting observation is the difference in behavior of a radial and
bias ply tire when the tire inflation pressure is zero or nearly zero.
Still ridial Spud
175 HR U/75 120 km/h
Pi • 1.8 bar 60 km/h
mm
8 mm tread dtpth
315

310
:3*5^~^ ^•^_
1 305
^**-
=-— R*
§.
300 ^^^ s'5?^^^
2M

290

650 1550 2450 3350 N


Tin load F,

FIGURE 8.2.43b Effective and loaded radius for a 175 UK- 14 radial tire.
584 MECHANICS OF PNEUMATIC TIRES

SUel radial Tread depth


175 HR U m
V • 120 km/h
p(. 1,8 bar

IMS J^C •^== Re


S.

"^ X^ ^\~
2M
"^

290
^^_ )Rl

(SO 1550 2450 3350 N


Tin load F,

FIGURE 8.2.44. For zero load it may be assumed R = R, = /?,.

Figure 8.2.45 and 8.2.46 show continuous measurement of loaded


radius during runs at several speeds, while simultaneously measuring the
tire load.
C. Effect of speed
The effect of speed has already been demonstrated in fig. 8.2.43a/b for a
bias tire and steel radial tire, as measured on a drum of 2.5 m in diameter.
Due to larger centrifugal forces at higher speeds the bias tire will grow, re
sulting in an increase of effective radius R,, increase of loaded radius R1
and an increase of free radius R. This effect is much less for the steel radial
tire. Both effective and loaded radii were increased slightly by an increase
in speed as shown in fig. 8.2.43b. Several authors report similar measure
ments. [31c]
D. Effect of braking and driving.
The effect of longitudinal loading on the loaded radius through braking or
traction forces is illustrated in figure 8.2.47 obtained on the Calspan Tire
Research Facility for radial and bias belted H78-15 tires [48a].

8.2.4 Rolling Resistance


Introduction
The rolling resistance of a free rolling tire is mainly caused by the inter
nal friction in the rubber and cord, while the slip in the contact zone and
the windage losses at moderate speeds are of less importance. Slippage of
tread surface on the road is about 5-10 percent of the total losses, while
the drag due to air friction represents 1-3 percent of the total loss. Other
MEASUREMENT OF TIRE PROPERTIES 585

Inflation pressure

0.0 bar
mm
,. 0.2 „
330
l
320

310

300

210

280
2
SL. 270

260

250

au
1000 1500 2000 2500 3000 N
Tire load FZ

FIGURE 8.2.45 Loaded radius vs. tire load at various speeds and inflations for a 175 HR-14
tire.

Inflation pressure
__.__._. — p, 0.0 bar
Bill ply
0.1 .
175 Sit
„ 0.2 ..

320
speed
310

g 280
•o

"- 270
•O

I 260
250.

1000 1500 2000 2500 3000 N


Tire load F2

FIGURE 8.2.46 Loaded radius vs. tire load at various speeds and inflationsfor a 175 S-14 bias
tire.
586 MECHANICS OF PNEUMATIC TIRES

TIRE: Radiol HR78-I5


_ INFLATION PRESSURE: 24 psl
ROAD SPEED: 30 mph
SLIP ANGLE: 0 degrees
VERTICAL LOAD
13.0 ''900 Ib —

•12.8
in
I 126

8 124

122

-1.0 -0.8 -0.6 -0.4 -0.2 0 0.2


LONGITUDINAL SLIP, s

TIRE: Bios-Belted
INFLATION PRESSURE: 24 psi
14.0 -SLIP ANGLE: 0 degrees

-1.0 -0.8 -0.6 -0.4 -Q2 0 0.2 0.4


LONGITUDINAL SLIP, s

FIGURE 8.2.47 Influence of longitudinal slip on tire loaded radius for both radial and bias-
belted tires.

factors influencing the rolling resistance are bad road conditions, in


volving large tire deflections, the presence of snow or the deformation of
soft soil.
At constant speed, a free rolling wheel requires a horizontal force Fr in
the wheel center to overcome the rolling resistance. The balance of mo
ments around the wheel center is shown in figure 8.2.48
MEASUREMENT OF TIRE PROPERTIES 587

TIRE : RADIAL NR7H5


INFLATION PRESSURE • I« pu
ROAD SPEED : 30 r,p*
SLIP ANGLE : 0

TIM : WAS-8ELTEDH7IIS
INFLATION PRESSURE : 14 (Mi
SLIP ANGLE : 0

FIGURE 8.2.47 (Con't)

where FN = load carried by the wheel h = axle height above the ground.
The rolling resistance force Fr is the resultant of the longitudinal tan
gential stresses in the contact patch, while the resultant /•'.. of the normal
force distribution has an offset, ahead of the contact center as in figure
8.2.48. The rolling resistance is commonly expressed per unit of load of
the tire, thus lh./ 1000 Ib. or kgf./1000 kgf. The unit of load varies in some
cases, such as earthmover tires, where it is denned as the resistance per ton
588 MECHANICS OF PNEUMATIC TIRES

Direction of motion

•Rolling resistance

FIGURE 8.2.48. Rolling resistance of a free rolling tire.

of load or for bicycle tires where it is given in lb./100 lb. or kgf./100 kgf.
The coefficient of rolling resistance fr is defined as
F F
f, = -=r or, expressed as a percentage, /, = -^- X 100%.

The rolling resistance of a tire is dependent on the load, inflation pres


sure, temperature, road conditions and speed. The level of the rolling re
sistance ranges from 10 to 25 kgf./ 1000 kgf. for passenger sizes. For ex
ample the rolling resistance of steel belted radial tires at a speed of 100
km/h may be (1.3-1.6)% of Fn, whereas at a speed of 180 km/h the range
is (1.9-2.4)% of Fn as measured on a steel drum. For commercial sizes
these values vary from (0.5-1.5)% of Fn.
From the above it follows that the rolling resistance of a free rolling
pneumatic tire is generally considered a force Fr rather than energy E, al
though this force is mainly the consequence of hysteretic energy losses ΔE
within the tire and contact area during travel. Because in the S.A.E. defi
nition (cf. ref. [3] sec. 8.3), the rolling resistance becomes negative at larger
braking torques it may be better, in order to avoid these inconsistences to
express the rolling losses of a tire in terms of energy. It is shown [49b] that
the energy loss

AL

per unit distance travelled. ΔL is always positive and therefore the energy
concept may be more meaningful than the force concept.
Methods of Measuring Rolling Resistance
A. On the road
Several methods are used to measure the rolling resistance on the road
under exact working conditions [49a].
a. Towing a vehicle at constant speed.
b. Allowing a vehicle to coast and measure the rate of deceleration.
c. Measurement of the torque in the drive shaft of the vehicle required
to maintain its speed.
MEASUREMENT OF TIRE PROPERTIES 589

TIRE MO-I1
kg)
p, . U k|l/eml

it I ' »
\CONTACT BEGINS CONTACT LENGTH cm
T
TIRE 5.90-13 FREE ROLLING
F,, -300 kgf
p, -15 kgf/cm2

FIGURE 8.2.49. The vertical force and longitudinal force distribution over the contact length
of a free rolling tire.

It has been found extremely difficult to find the rolling resistance accu
rately from these tests because it is not easy to separate other friction from
the measured results.
Rolling resistance trailers designed for towing have been used particu
larly to examine the effect of different road surface textures under working
conditions [49c].
The principal method of towing a trailer has been refined by towing a
shrouded passenger car over a range of speeds, and measurement of the
forces between the shroud and the enclosed vehicle (fig. 8.2.50a, b). This
approach, due to the Motor Industries Research Association of Great Brit
ain, involved the isolation of a complete car from aerodynamic forces by a
plywood enclosure mounted on a two-wheeled light weight trailer with
sheet rubber skirting attached to the bottom of all the sides.
A similar arrangement is the towing of a shrouded single wheeled trailer
at constant speed, and measuring the towing force by means of a dyna
mometer.
Because the horizontal force to be measured is only about 1 percent of
the vertical load on the trailer, the determination of differences due to load
changes, air pressure, or tire construction characteristics calls for an ex-
590 MECHANICS OF PNEUMATIC TIRES

(a) Vehicle rolling resistance trailer

LATERAL POSITIONING LINKS PLYWOOD BOX

TOWING VEHICLE

(6) Vehicle rolling resistance trailer

FIGURE 8.2.50. Method of measuring the rolling resistance by lowing a shrouded passenger
car.

tremely high accuracy of measurement and carefully controlled conditions


of operation [50].
Difficulty has been experienced in obtaining a constant pull at constant
speed, because of the inertia of the trailer, but results are also dependent
on the gradient of the test road surface, the friction of the wheel bearings,
climatic conditions, etc. It is almost impossible to control the tire temper
ature. In view of all these difficulties, it is generally accepted that indoor
machine tests give more reliable comparative values between different
tires or operating conditions.
B. Indoor Machine Tests
Practically all methods measure the rolling resistance by loading the tire
against a smooth-faced steel drum, and by accepting that significant dif
ferences in tire behavior are induced by the curvature of the drum so that
the values obtained are not those which would be obtained under actual
road conditions.
Due to curvature effects it appears that a moving flat surface is the best
for laboratory measurements of rolling resistance. Movable platform ma
chines are not used for this purpose because it is not possible to reproduce
correct temperature effects in the tire. The accuracy and precision of the
moving belt machine developed by the Calspan Corporation makes it pos
MEASUREMENT OF TIRE PROPERTIES 591

sible to identify most tire parameters that may have a bearing on the en
ergy consumption.
The obvious advantages are that it is comparatively simple to maintain
control over the variables speed, load, inflation pressure and temperature.
Four methods are used to determine the rolling resistance under free
rolling conditions on drums:
a. In the inertia method the tire is loaded against a drum of known mo
ment of inertia and the horsepower consumption of the tire is found
from the rate of deceleration of the drum through the chosen speed or
range of speed. The test is repeated with the tire in skimming contact
under zero load and the drum slows under bearing friction and the
windage of the drum and tire. From both speed-time curves the rolling
resistance is determined. This simple method has the disadvantage that
it is impossible to determine the rolling resistance under stationary
conditions at a given speed, and only limited control of temperature is
possible [49-51].
b. In the torque-shaft method the torque M can be determined on the shaft
in the drive to the drum by measuring the twist of a torsion shaft. The
twist is measured both with the tire in skimming contact and with the
tire loaded against the drum. The difference between the torques is
used to determine the rolling resistance or the coefficient of rolling re
sistance /„ and can usually be measured with an accuracy of 0.5-1.0
percent [52].
c. In the motor torque reaction method the motor driving the drum is trun
nion mounted in order to measure the torque reaction in the stator by a
rigidly attached moment arm which engages with a spring balance or a
load cell [51]. The measuring technique is the same as in test (b), and in
both methods constant speed is very important to avoid errors due to
acceleration and deceleration of high inertia test drums.
d. The tire to be tested with the six-component tire tester is mounted in a
rig with a loading device above the drum (fig. 8.2.51). This method
calls for an extremely high accuracy in placing the center of the tire in
a vertical direction above the centerline of the drum. It is convenient to
average results obtained from tests with clockwise and counterclock
wise rotations of the drum.
Tire Conditioning
Since the rolling resistance of a new tire decreases during the first hours
of running, the tires are preconditioned by running them 2-15 hours at
about 50 km/hr. under 80-100 percent of their maximum scheduled load
and inflation pressure. The tire will deform from its cured state to an equi
librium condition by adjusting its localized stresses internally. A distance
of 150-300 km. is, however, often considered sufficient as a break-in pe
riod to bring the tire down to its ambient level of power loss [53a and 53b].
After preconditioning, the load, speed, and inflation pressure are ad
justed to the desired values for test, and the tire is run until the temper
ature of the air contained within the tire remains constant for at least 10
minutes. Readings of rolling resistance and the temperatures of the con
tained air and the ambient air are recorded. Immediately after readings
are taken, the tire is lifted from the drum and stopped. Then the tire tem
perature is measured by a thermocouple needle at about four preselected
points in the shoulder and in the tread.
The sequence of measuring the various combinations of load, speed,
592 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.51. Mounting of the six-component Delft tire tester in a rig with an air-spring
loading device above the drum.

and inflation are varied in accordance with a statistical design, in order to


eliminate systematic bias [54].
Parameters Affecting the Power Loss
The most important factor influencing the power consumption in the
rolling tire is the hysteresis of the materials, representing 90-95 percent of
the total power loss. Decrease of the deformation of the tire during rota
tion through the contact zone and decrease of vibrations after deformation
will help to minimize the power loss, as well as optimizing the tire struc
ture and rubber compound to lower hysteresis loss [53].
A. Effect of tire load, radial deformation and inflation pressure.
An increase in load at constant pressure and speed results in larger tire
deflection and consequent increase in rolling resistance as illustrated in
figure 8.2.52.
An increase in inflation pressure at constant load and speed results in
less tire deflection, lower hysteresis loss and consequently, less rolling re
sistance [55]. It also may reduce the amount of scrub because it reduces
the size of the contact area.
B. Effect of speed.
The effect of speed is shown in figure 8.2.53 with constant load and con
stant inflation pressure [51]. The data are plotted in the form of rolling re
sistance against speed as curves of constant temperature. At higher speeds
a progressive increase is observed as indicated in the lower figure. At con
stant temperature curve A indicates the behavior, while curve B shows the
MEASUREMENT OF TIRE PROPERTIES 593

COEFFICIENT ROLLING RESISTANCE % OF F.

1.0 u
1 cm. J.35 1.50 1.61
V - X km/h 2.40 1.91 1.61
2 cm.
165-13 2.46 1.83
3.51 2.12 172

11
INFLATION PRESSURE Pi kjt/crn'
»
1.5
1.0
i 10

100 200 300 400 500 600 700 NO 900

FIGURE 8.2.52. Effect of vertical tire load and inflation pressure on the rolling resistance.

effect of the higher "service" temperatures. Because the inflation pressure


was kept constant during these tests, there is an additional stabilizing ef
fect tending to further reduce the increase of rolling resistance with speed,
as a result of the rise in inflation pressure as the tire gets hotter. There is
another "fortunate" effect in that the hysteresis loss decreases as the tem
perature increases. Also, slippage losses in the contact area are reduced at
higher tread temperature due to a lower coefficient of friction.
When starting at room temperature conditions, the rate at which the
temperature of a passenger tire rises and the rolling resistance diminishes
is shown in figure 8.2.54. Both level off to constant values. The curves rep
resent constant load, speed and inflation pressure.

§ Tirt 7.50.U tubtltss


{
E
2- F..tlOkgf
S P: -1.7kaf/cm2(conftinO

30 At conit»nt ttmp«rjtur«

25

20 tif v let ttmp«ratuft

SO 60 70 10 90 100
••km/h 100 » 125
Spttd150
km/h

FIGURE 8.2.53. Effect of speed on the rolling resistance.


594 MECHANICS OF PNEUMATIC TIRES

ROLLING RESISTANCE kjt TEMPERATURE IN SHOULDER


IM
| DRUM C
T TIRE 72
_TEMPERATURE_ c V >I(S km/h

I XX'
1 F.-400

20
/
It
|N^_ •M

f,

0 10 20 30 TIME (min)
EFFECT OF TEMPERATURE ON ROLLINS RESISTANCE

FIGURE 8.2.54. Effect of temperature on rolling resistance.

The equilibrium temperature as measured with a thermocouple needle


in the shoulder after a run of 30 minutes is shown in the form of temper
ature against speed in figure 8.2.55. On flat roads these temperatures will
be lower due to less deformation. The temperature of the contained air is
about 25-30°C lower than the shoulder temperature at equilibrium condi
tions.
C. The effect of rubber compound and cord.
The hysteresis losses within rubber and cord are well known. It follows
that the higher the hysteresis losses the higher the rolling resistance. This
subject has been treated in chapter 1. Readers are also referred to details
given about energy losses due to tread, side wall, ply rubber and cord in
the literature [53, 55, 56a, 56b].
D. The effect of carcass construction.
"The difference in behaviour for three types of construction, Radial,
Bias belted, and Cross ply, is shown in Fig. 8.2.56. As can be seen, the
radial ply construction, having an inextensible belt, shows an advantage
up to speeds of 145 km/h (90 mile/h). It is well known that the reduction
in rolling resistance on passenger car tires worn to the wear indicators is
about 15 percent."
TEMPERATURE IN SHOULDER

DRUM DIAMETER 2.5m


A

^s
,—
F,, -400 kgf
Pi- .5 kjf/em*

10- .^
*-^^
•—
7
• SO 100 ISO MO SPEED km/h
EFFECT OF SPEED ON EQUILIBRIUM TEMPERATURE

FIGURE 8.2.55. Effect of speed on equilibrium temperature in shoulder.


MEASUREMENT OF TIRE PROPERTIES 595

0.03

Crossply

Bias belted

Radial Peterson
0.01

40 80 120 160(km/h)

20 40 60 80 100(mile/h)
Speed
FIGURE 8.2.56 Effect of speed on the rolling resistance coefficient of three different types of
tires.

Changes in the design parameters, such as the angle of the cords at the
crown, the number of plies in the carcass and the rim width affect the de
formation of the tire and consequently the rolling resistance. Truck tires
appear to have lower rolling resistance values (5-10 kgf./1000 kgf.) than
passenger car tires, because of other tread rubber compounds, higher in
flation pressure and radial stiffness [54].
Lower cord angle increases the radial stiffness of the tire and con
sequently decreases the deformation at a given load and inflation. Because
the circumferential stiffness is also increased due to a lower cord angle the
interia losses are also decreased at high speed [58].
Changing from four-ply to two-ply design gives a reduction in the hys
teresis losses which is related to the mass of the tire and to an increase in
the deformation. The combined effect of these two factors leads to a de
crease in the rolling resistance [58].
The effect of the rim width is that mounting of a conventional bias tire
on a wider rim gives a decreasing tire deflection, and therefore a reduction
of the rolling resistance [51, 58].
The height-to-width ratio, known as the aspect ratio of the tire, has re
cently received much attention. The trend towards lower aspect ratios will
continue, and low ratios of about 0.50 are already installed on cars, com
pared with 0.84-0.95 for bias tires of several years ago [59].
Lowering the aspect ratio is advantageous with respect to rolling resis
tance because the radial stiffness is increased and the tread is flatter, which
also decreases the deformation and consequently the hysteresis loss. Due
596 MECHANICS OF PNEUMATIC TIRES

to an increase in circumferential stiffness the inertia losses at high speed


are reduced.
At low speed the section height effect is greater for the bias tire than the
radial belted, but the high speed performance shows significant improve
ments for the radial belted tire. [49c, 56b]
Cornering
As the slip angle increases the rolling resistance increases due to a com
ponent of the cornering force Fy in the direction of motion, and also due to
a slight increase of Fr as shown in figure 8.2.57.
, Braking and Traction
The effects of braking and traction are shown in figure 8.2.58, where the
rolling resistance values are plotted against braking and traction coeffi
cients. These coefficients have been defined as the tangential braking or
traction force FBT divided by the radial tire load FN. Due to an increase in
tractive or braking effort an increase in longitudinal slip is observed, re
sulting in a higher rolling resistance [51, 60].
The rolling losses under braking and traction conditions are equal to:

FB V - A/rfi - FTV
respectively, where
FB = - F, = braking force
FT = + F, = traction force
A/g = braking torque
MT = traction torque
il = angular velocity
V = speed.
Figure 8.2.58b shows eR for a 6.45S14 tire at zero slip angle and it appears
that the minimum energy loss occurs when the tire is driven and not under
freely rolling conditions. Schuring [49b] gives an explanation that the
driving torque, when small, would reduce slippage losses (indicated by the

T- .« ,/ IN DIRECTION OF MOTION
Tire 7. SO -1* t, eof«.F.«inw
F. .100 kjf ./ '
fi .1.7 kf l/e m« X
u
JO /

/
IN WHEEL PUNE
10 I^M^KlWww*^** Fr

• • " 1
0* '_ '.
SLIP ANGLE a
FIGURE 8.2.57. Effect of slip angle on rolling resistance.
MEASUREMENT OF TIRE PROPERTIES 597

MOILING RESISTANCE
Tirt t.n-K
F..-IW kof

^ 0.4 O.J O.Z 0.1 0.1 02 0.3 0.4 .


BRAKING COEFFICIENT DRIVING COEFFICIENT

FIGURE 8.2.58a. Effect of braking and traction on rolling resistance.

negative slip ratio at and around the freely rolling condition) more sharply
than build up additional deformation losses.
Temperature Measurements
Because tire temperatures are related to power and affect the service life
of tires, the measurement of the tire temperatures is considered as a re
liable indicator of the energy put into it and the durability of the tire.
Due to deformation and hysteresis, heat is generated within the tire.
Studying the heat build-up has always been an important factor in assess
ing the durability of the tire. This is because severe testing is often associ
ated with tread separations from the cord body resulting from thermal
degradation and high centrifugal force.
Operating temperatures have been measured while the tire is running
by means of thermocouples vulcanized or inserted into the tread or plies,
but the life of these thermocouples is short because of fatigue breakage
under load.

-500 500 1000


E.lb

FIGURE 8.2.58b. Effect of longitudinal force Ff and normalforce F, on energy loss eK for a
6.45S14 tire at zero slip angle.
598 MECHANICS OF PNEUMATIC TIRES

More frequently, measurements have been made by inserting thermo


couple needles under the tread or in the plies after stopping the tire. Mea
surements of the temperature of the air in a tire are also made, as well as
determining the tire surface temperature.
Figure 8.2.59 shows typical values of tire temperatures taken on an in
door test wheel as a function of speed, and with various test conditions. It
appears that many normal passenger car tires operate in the vicinity of
100°C at normal conditions, measured in the air cavity [62].
Several authors report the measurement of tire surface temperatures us
ing a contact thermometer or thermistor to relate the average skin or
groove temperature to heat loss or the relative wear rating of tread com
pounds [61, 56].
Other methods were developed for determining the temperature of the
air in a tire [63].
The difficulties encountered using thermocouple needle measurements
are also reported in the literature [56, 64, 65).
Refinements in measuring techniques made it possible to measure inter
nal tire shoulder temperatures under road testing conditions of high load
ing and speeds, to study durability and reliability of the tire [65, 66]. In en
durance testing on indoor machines it has often been found advantageous
to obtain the hot spots before severe tread separation occurs. The use of
lead-sulphide photo-conductive cells for high speed pyrometry has been
reported [67, 68, 69J.
Truck tire rolling resistance
Compared with normal passenger car tires, trucks or other commercial
vehicle tires have load carrying requirements that are much higher, up to
10 times as much.
Generally this kind of tire requires a large number of plies and has in
flation pressures 3 to 4 times higher than car tires. This occurs because the
deflection at a fixed load is less at higher pressures, thus reducing the hys
teresis loss which is dependent on the tire deformation. Due to deforma-

COMPOS.TEOATA TEST CONDITIONS.


11.040 RANGE 6 TIRES) «O*F AMBIENT CCMPOSJTE DAT* TEST CONDITIONS
8 23-)4 BIAS 28 PSi (COLD) INITIAL (LOAD RANCE B TIRES) 120' WHEEL
C 78 -14 BELTED BIAS PRESS 8 23-14 BIAS >OO*F AMBIENT
G 70 -14 BELTED BIAS LOAD- O8O LBS 078-14 BELTED BIAS
?05fi Ml RADIAL GENERAL TIRE UVALDE 0 70- 14 BELTED BIAS PRESS
4Q TRACK 2O3RI4 RADIAL LOAD-rJdOLBS
EQUIUILENT
(I2BSLU ON I2O'

•— ,——
—•— —•—

*. 30 -, p
2

SHOULDER
^—- ^ CONTAINED
«-^"*l
——• AIR
==
too 4^ (a)
" 60 70 80 90 IOO GO 70 80 90
SPEED MPH SPEED MPH

FIGURE 8.2.59. (a and b) Equilibrium pressure and temperature rise as functions of tire load
and speed.
MEASUREMENT OF TIRE PROPERTIES 599

1400'LOAD
•1200* LOAD
OOO'LOAD

(c)
40 50 60 TO 80 90 100
SPEED MPH

FIGURE 8.2.59 (c) Equilibrium pressure and temperature rise as Junctions of tire load and
speed.

tion, heat is generated throughout the carcass and tread region by cyclic-
stress and material hysteresis losses. This heat build-up is a serious prob
lem for truck tires and therefore natural rubber components having low
hysteresis characteristics are used.
Another way to improve the heat build-up through tire construction
and design is the application of steel wire cord. Recently there is a trend to
produce all-steel radial truck tires, having a smaller number of plies than
designs with other cord materials. The dimensions of the beads are re
duced as well, because of the reduced number of plies in the carcass.
These factors opened the possibility of producing tubeless truck tires. Be
cause these tires can be mounted and dismounted like car tires, there is a
tendency to use more tubeless truck tires.
The rolling resistance of truck tires, as a percentage of the vertical load,
is low compared with car tires. The level of rolling resistance ranges from
0.5 to 1.2% of Fz, while car tires are in the range of 1.3-1.8% of Fz at 100
km/h. The lowest value of the rolling resistance for truck tires is found for
a so called "super single" tire, which replaces dual tires as shown in figure
8.2.59d.
Radial truck tires applied as dual tires are very sensitive to differences
in the circumference. A small circumferential difference causes high forces
and extra longitudinal slip in the contact area, and therefore increases the
tread wear. These effects for the free rolling dual truck tires are shown in
figure 8.2.59e.

8.2.5. Braking and Traction


Braking
A. Deformations
The difference in deformations between a free rolling and a braked tire
has already been discussed in chapter 6. However, to facilitate the under
600 MECHANICS OF PNEUMATIC TIRES

40 rolling resistance
coefficients:
dual 0.9% of Fz
30 wide 0.6% of FZ

S
S 20 =0

0 1000 2000 3000 4000 5000


Verticil load Fz ( kgf )

FIGURE 8.2.59. (d) Rolling resistance for dual pair vs. wide base single truck tire.

standing of the following paragraph, another illustration of tire deforma


tions shall be given [24].
To show this difference in behavior, a photograph was first taken of a
freely rolling tire and afterwards, using the same negative, a second photo
on top of the first one has been made of a braked tire. Triggering of the
camera has occurred at exactly the same position as in the free rolling
case. This is shown in figure 8.2.60.
By measuring the distance between the radii which were originally
equally spaced on the surface of the undeflected tire, the deformations of
tread and sidewalls can be determined for various braking conditions.
Figure 8.2.61 shows some results thus obtained on the movable platform
machine, and it is seen when a braking force is applied that the tread ele
ments before contact are stretched. The alterations of the deformation are
also clearly shown on a photo of the contact area of a braking tire in figure
8.2.62.

DUAL TIRES
10.00-20 UPR
free rolling

-no

-200

Effect of circumferential difference for dual tires.

FIGURE 8.2.59. (e) Effects of size difference on dual pairs of truck tires.
MEASUREMENT OF TIRE PROPERTIES 601

FIGURE 8.2.60. The difference in deformation between a free rolling and a braked tire.

B. Force distributions
a. Longitudinal force distribution
The movable platform machine is also equipped with an apparatus in
corporating three measuring bars 2 cm. long and 20 cm. wide for measur
ing the distribution of the vertical, longitudinal, and lateral forces [14, 24].
With this modified "Gough" apparatus shown in figure 8.2.63, the dis
tribution of the longitudinal shear forces in the contact area is determined.
The distribution of the forces created by a free rolling tire is represented
by line 1 in figure 8.2.64, and the additional shear force created by the
braking torque is represented by line 2. The resultant shear force distribu
tion along the length of contact is therefore represented by line 3, as mea
sured with the longitudinal force bar of the Gough apparatus. The precise
form of this curve depends very much on the magnitude of the braking
force for a given radial load, inflation pressure, coefficient of friction, etc.
as described in chapter 5, figure 5.52. The reason for such a force distribu
tion becomes clear by looking again at figure 8.2.61. An extended tread
element adheres to the platform on first entering the contact zone (ch. 6,
fig. 6.1c). As it moves further into the contact area, it produces a deflection
which increases linearly with increasing distance (causing an increasing
longitudinal force) until the local value of limiting frictional force is
reached and the tread element begins to slide back, thus reducing again
the longitudinal force as shown by line 3 in figure 8.2.64.
The total longitudinal force /•'„ may be obtained using the six-com
ponent tire tester, also mounted in the Delft movable platform machine of
figure 8.2.5. The result obtained can be compared with the value obtained
from the longitudinal force measuring bar, by integrating the longitudinal
force function as the tire travels over it.
602 MECHANICS OF PNEUMATIC TIRES

Braking force Q

Tire S.90-13
Bias carcass s '.
1/ -
Braking force 200 kgff. \
Load 300 kgf
In f(. pressure 1.5 kgf/ cm*
r' i
/
braking v "7
undetected
•'

FIGURE 8.2.61. Deformations due to braking.

FIGURE 8.2.62. Photo through glass plate of contact area of braking tire.
MEASUREMENT OF TIRE PROPERTIES 603

FIGURE 8.2.63. Modified Gough apparatus incorporating three bars measuring the lateral,
longitudinal and vertical partial forces.

b. Vertical force distribution


The vertical force distribution is measured by a second bar of the
Gough apparatus. The difference in behavior of a free rolling and braked
tire is shown in figure 8.2.65. The total vertical force FN may be obtained
either by integrating the vertical force function, or by using the tire tester,
the latter method being preferable. The offset of FN before the contact cen
ter may also be obtained. As already discussed in section 8.2.4, the coeffi
cient of rolling resistance fr is higher than in the free rolling case, mainly
due to a rise in slip-induced rolling resistance. The balance of moments
around the wheel center 0 reads (fig. 8.2.64):
FB • h = MB + F»f,.
Due to the braking torque, it is observed that the distance // is smaller than
in the free rolling case [24].
c. Longitudinal sliding
Attempts have been made to measure the longitudinal sliding with a
small cam wheel mounted in the longitudinal force bar of the Gough ap
paratus. The cam wheel can rotate, but does not measure the longitudinal
sliding of one single tread element because due to the sliding motion the
neighboring tread element subsequently comes into contact with the cam
wheel. The total longitudinal sliding distance is measured with a poten
tiometer coupled to the cam wheel shown in figure 8.2.63. It is the sum of
the longitudinal sliding of all different tread elements along a line parallel
to the direction of travel of the platform. These measurements of sliding in
the contact patch were not accurate enough. Photography of the contact
604 MECHANICS OF PNEUMATIC TIRES

Vo A

Platform velocity

Vertical force
distribution

Longitudinal force
distribution

Longitudinal
sliding distance

I * sliding velocity

VO-O.R

Velocity

Tire circumference

FIGURE 8.2.64. Distribution offerees, longitudinal sliding distance and sliding velocity, over
the contact length of a braked tire.
MEASUREMENT OF TIRE PROPERTIES 605

Tire 530-13 bit* ctrcait


FN i 300 kgf
FB , 200 kgf
E SO Pi i 1.$ kgf/cm1
_ tree rolling
*0 bra k i n g

30

I »
^

S 10 20
^Contact begint
Contact length cm

FIGURE 8.2.65. Comparison of vertical force distributions over the contact length of a free
rolling and braking tire.

area of the tire, marked with a grid, during motion of the glass surface of
the platform machine, appeared to give better results (fig. 8.2.62).
Due to stretching of the tread elements before contact the circum
ferential velocity will increase, and on the assumption of no sliding in the
front part of the contact zone for a moderate braking force, the tread ele
ments coming into contact with the platform will start to travel with the
platform speed V. The longitudinal shear force which increases towards
the rear of the contact zone, in combination with the decreasing vertical
force, will cause rearwards sliding of the tread elements in the rear part.
The resulting longitudinal sliding distance and sliding velocity curves,
taken over the contact length, are shown in figure 8.2.64. Increasing brak
ing force at constant vertical load will result in increasing sliding over the
contact length as in figure 8.2.66.
The slip ratio may be defined to be:

where J20 = angular velocity at free rolling


fis = angular velocity at braking
Both values can be measured at contant forward speed of the platform or
at constant drum speed [sec. 8.2.3, fig. 8.2.42]. As the braking force FB in
creases so does the percent slip. The resulting braking force coeffi-
p
cient -rr- is shown in figure 8.2.66 for constant speed as measured on a dry
fN

steel drum of 2.5 m. in diameter.


606 MECHANICS OF PNEUMATIC TIRES

free
•raking force ,0[lmg
coefficient wheel

I eoniUnt
1 speed

brsking slip percent K


B Wo

sliding

wheelsllp S'/. WY. IS0/.

FIGURE 8.2.66. The relation between braking force coefficient and braking percent slip.

The difference in behavior of a radial ply and bias ply tire is clearly
demonstrated in figure 8.2.67 at a constant speed of 40 km/hr. [70]. It is
seen that the position of the peak coefficient μlp is often very difficult to de
termine since the curves are sometimes rather flat. Similar results are also
often obtained on wet surfaces [71].
D. Effective radius
Returning to figures 8.2.61 and 8.2.62, the stretch of the tread elements
just before contact, as measured from the photograph, appears to be ap
proximately 10 percent. Assuming again no sliding in the front part of the
contact zone, the same reasoning as in the free rolling case (sec. 8.2.3) re
sults in: V = 1.1Vo, that is Re = I.IR, showing an increase in effective
radius compared with the free rolling condition (Re = !).%/?) and a de
crease in angular velocity fl [70].
The slip ratio may be defined to be:

*„ =
Q.
where /?,„ = effective radius at free rolling
Rea = effective radius at braking.

E. _Wet_road measurements
a. Distance method [72]
MEASUREMENT OF TIRE PROPERTIES 607

The locked wheel "sliding" friction coefficient is determined during the


speed interval under investigation.
The calculation of this friction value, sometimes called "braking coeffi
cient" is based on the accurate measurement of the skid distance s and the
speed V. It follows that:
V\-V\
2gs

where V\ — initial velocity


K2 = final velocity
g = gravitational constant.
Due to the erroneous assumption in the above formula of a constant de
celeration over the speed interval (ju being speed dependent) it is rather
difficult to compare results obtained with different tires on various road
surfaces. Although standard vehicles are used, a large number of tests are
required to arrive at an acceptable accuracy of the averaged ju,, values.
This is usually done at speed increments of about 10 km/hr. in the re
quired speed range of the test vehicle.
b. Deceleration method
As discussed in chapter 6, the braking performance of tires can also be
measured by the deceleration of the vehicle. In this method only the front
wheels of the test vehicle are braked in order to maintain directional sta
bility of the vehicle at all speeds [73].
Both the peak value of the braking force coefficient /i,,,, and the locked

Onim mtiiurtmtnt
V.iOkm/h
F<=100hjf

Slip

FIGURE 8.2.67. Comparison of the braking force coefficient-braking slip relationships of a


radial ply and bias ply tire at constant drum speed.
608 MECHANICS OF PNEUMATIC TIRES

wheel slide value μls are determined from the film record of deceleration
(ch. 6, fig. 6.11).
Because these tests can be executed on a relatively short test surface,
having a more or less constant factional character and uniform water
depth, comparison of tire data thus obtained is usually preferred over the
stopping distance method. The distance method with its long skid dis
tances has a poorer accuracy, especially in the higher speed range [72].
It has often been observed with the deceleration method that radial ply
tires gave higher peak braking force coefficients filp on all surfaces than do
bias ply tires. The differences between these peak coefficients were almost
independent of speed and were least on coarse-textured surfaces, and
greatest on fine-textured surfaces, ranging from almost zero to about 0. 1 .
The radial ply tires gave higher locked braking force coefficients ^ on the
fine textured surfaces but lower coefficients on the coarse textured surfaces
than did the bias ply tires [73].
c. Force method
Numerous braking force tests with different tire constructions and tread
rubber compounds have been executed on a variety of dry and wet road
surfaces and are reported in the literature.
The advantages of the towed trailer road tests are that the towing truck
can maintain constant speed at the desired level, uniform wetting immedi
ately ahead of the tires is obtained, and a large variation in tire loading is
possible without any load transfer effects (fig. 8.2.8).
During these road tests it appears very difficult to obtain braking force
coefficient values in the range between the peak braking force coefficient
H,p and the locked braking force coefficient /i,, (κb - 100 percent), because
of almost immediately locking of the wheel after ^ has been reached. For
this reason, usually only ju,, and /i/% are determined to compare tread de
signs, rubber compounds, the constructions etc. The numerous results ob
tained are very similar to data published in the literature [73] and dis
cussed in chapter 6.
Another method to determine the braking force coefficient has been
published by the Road Research Laboratory [71]. Force measurements are
made on a fifth wheel mounted in a test vehicle. The angular velocity of
this test wheel can be held at any desired value, independent of the vehicle
speed.
At the beginning of a series of tests a record of wheel rotation is ob
tained with the free rolling tire in order to obtain its rolling radius Rr., for
evaluation of the braking slip.

Typical curves of braking force coefficient against braking slip κB are


shown in figure 8.2.68 for a bias ply tire on three out of five of the test sur
faces from table 8.2.1.
The curves show in general a rapid initial rise, and then become gradu
ally less steep as they approach a peak at 7-15 percent slip. Beyond the
peak there may be a slowly tailing region and finally a more rapid fall
from 80 or 90 percent to the locked wheel values at 100 percent slip.
The rough and harsh surface No. 1 is capable of giving a high coeffi
cient. The harsh microtexture has the largest effect on the coefficient in
MEASUREMENT OF TIRE PROPERTIES 609

Tin bin ply 125.16


F^.230 kgf
Pi-Vl kflf/cm*

20 30 <0 SO SO
Braking slip per ctnt

FIGURE 8.2.68. Typical curves of braking force coefficient against braking percent slip for a
bias ply tire on three road surfaces at two speeds.
Surface I (rough, banh). surface 4 (rough, polished), surtax S (smooth, polished).

comparison with the other surface No. 4. It has been shown that tires of
radial ply construction have a more rapid initial rise of the braking force
coefficient curve than bias ply tires. On harsh surfaces the radial ply tires
gave higher peak coefficients but on polished macadam surfaces lower val
ues were observed. The same was true but less obvious for the locked
wheel case [71].
Curves of pure braking force coefficient against braking slip give valu
able data for development of anti-skid braking systems. These systems are
of particular value when braking and cornering forces act simultaneously
[71, 74].
Traction
A. General observations
Having discussed the behavior of a braking tire at length in the previous
section, a short description of the action of a driving torque MT will be suf
ficient because the situation is analogous.
When a tractive force /', is applied, the tread elements ahead of contact
are compressed. The resultant shear force distribution along the length of

TABLE 8.2.1
No. Description Texture
1 9.5 mm. quartzite macadam Rough, harsh
carpet
2 Fine cold asphalt Smooth, harsh
3 9.5 mm. mixed aggregate Rough
macadam carpet
4 9.5 mm. Bridport macadam Rough, polished
carpet
5 Mastic asphalt Smooth, polished
610 MECHANICS OF PNEUMATIC TIRES

contact, as measured with the longitudinal force bar of the Gough appa
ratus, is shown in figure 8.2.69. As discussed (ch. 5, fig. 5.53), the form of
this curve depends on the magnitude of the tractive force. The resultant
shear force distribution (line 3) can be seen as the sum of the force distri
bution of a freely rolling tire (line 1) and the additional shear force created
by the tractive force (line 2).
A compressed tread element adheres to the moving platform when first
entering the contact zone. As it moves further the increasing deflection of
the tread element produces a linearly increasing longitudinal force. Begin
ning in the rear part of the contact zone, forward sliding of the tread ele
ment will be observed [24]. The resulting longitudinal sliding distance and
sliding velocity are also shown in figure 8.2.69.
The slip ratio may be defined to be:

where fl0 = angular velocity at free rolling


Qr = angular velocity with traction.
As the tractive force FT increases, so does the percent slip, and the result-
p
ing traction force coefficient ~rr-as a function of percent slip may be
FN
plotted, giving a curve similar to that obtained for the braking force coeffi
cient.
The decrease in effective radius ReT compared with the free rolling case
may be obtained by measuring ft, and ft,, at constant platform speed or
drum speed.
B. Wet road measurements
a. Break-away method
A light truck fitted with the test tire is attached to a large dynamometer
truck. The dynamometer truck is used for braking and maintains a low
constant forward speed of about 5 km/hr. to ensure rolling of the test tire.
By measuring the maximum tractive draw-bar force developed by the tire
in the low slip range of less than 15 percent, the test determines the maxi
mum force available to accelerate a vehicle, as the torque is gradually in
creased until the tire slides completely [72].
b. Dynamic slip method
This is a continuation of the break-away test, because dynamic testing is
conducted at 75 to 200 percent slip.
Traction ratings for both break-away and dynamic tests are given as a
ratio of tractive force to tire load. Typical recordings are shown in figure
8.2.70 [72].

c. Maximum acceleration method


The tire comparisons are based on the speed at wheel spin in the speed
range 40 to 100 km/hr. using a passenger car. The procedure is to drive
the car on the test surface at constant speed, then accelerate by means of a
full throttle down shift, or limiting the throttle position in order to reduce
MEASUREMENT OF TIRE PROPERTIES 611

I Platform velocity

sliding velocity

Tire circumference

FIGURE 8.2.69. Distribution offorces, sliding distance and sliding velocity over the contact
length of a tire under the action of a driving torque MT.
612 MECHANICS OF PNEUMATIC TIRES
-Breakaway

-Dynamic

50 100 150 200


SlipVo
FIGURE 8.2.70. Relation between relative traction force percent and traction percent slip.

the maximum available torque to a level compatible with the test condi
tions of surface and tires.
The driving force vs. speed curves of the different gear ranges for the
test car are usually first obtained on a drum dynamometer or derived from
engine torque curves. The theoretical point of slip is shown in figure 8.2.71
at the intersection of the maximum available driving force and the tractive
coefficient curves for two different tires on the same test surface. The speed
at the point of wheel spin indicates relative tire traction. Tire comparisons
are based on the maximum speed attained before wheel spin, or the least
time required to travel a given distance. For the latter test, slip may occur
throughout the test distance [72].
Several other methods are used to measure the tire resistance to wheel
spin under conditions of acceleration and with the vehicle travelling
straight ahead. Among these the single tire test technique is reported [72,
75].

first gear

0 20 60 100 140 180


Vehiclt spaed km/h

FIGURE 8.2.7 1 . Illustration of maximum acceleration method giving the relationship of tire
traction to drivingforce.
MEASUREMENT OF TIRE PROPERTIES 613

C. Factors affecting traction on wet roads


The tire variables affecting passenger tire traction on wet road surfces
include: tread design, tread compound and tire construction (ch. 6).
The tread design variables are bladed tread pattern versus plain design,
groove pattern, tread width, tread radius and center shoulder effect.
The tire construction variables such as crown angle, tread reinforce
ment, and radial construction versus bias construction have a smaller ef
fect on the wet tractive effort than the tread design variables.
Tread compounding is another complicated subject. For further infor
mation the reader is referred to the literature [59, 76].

8.2.6 Tire Nonuniformities


A. Dijks
Introduction
As road surfaces became smoother their contribution to vibration de
creases and increased attention has been focused on the tire-excited vibra
tions introduced by the nonuniform properties of the tire.
It is impossible to manufacture perfectly uniform tires because each
item has its own manufacturing tolerances. Only the rigid control of all
processes throughout the manufacture of materials and components for
tire building can minimize the unavoidable imperfections affecting uni
formity.
Moreover the production of a tire is characterized by a great amount of
handwork; deviations from the ideal construction are therefore inevitable.
Ply splices, non uniform fabric and steel properties, uneven curing, eccen
tricity of a belt mounting and other points of poor workmanship will pro
duce irregularities in the tire.
In the future it may be possible to produce tires by means of direct
molding [77a, 77b]. In this way a cordless tire consisting of only a few
components can be produced in a fully automatic procedure. These com
ponents are different types of rubber for the sidewall, tread, inner liner,
and steel wires in the beads. This method may result in very uniform tires.
Lack of uniformity around the tire will produce variations in forces ap
plied by the tire to the vehicle, and repeat its influence with each revolu
tion of the tire. The resulting periodic vehicle vibrations are speed depen
dent and often very annoying to the driver and passenger. Parts of the
vehicle vibrate and radiate energy as sound heard by the passenger, and
vibrations of the steering wheel, floor and seats are felt directly by the
driver and passenger.
The transmission of vibrations by the tire caused by road irregularities,
or by the natural frequencies inherent in the tire structure, are not consid
ered in this section, since these characteristics are not dependent on the
degree of nonuniformity. But it should be observed that chassis tuning ca
pable of attenuating all tire vibrations resulting from various road irregu
larities is extremely difficult, even more so because the various related vi
bratory systems should also avoid natural frequencies which might
coincide with natural tire frequencies.
A normal production tire will show nonuniformities that can be divided
into three groups:
A—dimensional nonuniformities
B—force variations for the free rolling tire, even at slow speed
C—mass unbalance
614 MECHANICS OF PNEUMATIC TIRES

lamp

optical
transducer
screen tire

10 km/h 180 km/h


JRevolution,
Radial runout
— Mean value
IRevoLution Growth of the
outer diameter
of the tire

FIGURE 8.2.72. Radial run out measurement and an example of results for different speeds.

A. Dimensional nonuniformities
The dimensional nonuniformity of a tire is usually characterized by
radial and lateral run out. The radial run out of an unloaded free running
tire is defined as the variation of the radius of the tire measured in a plane
perpendicular to the spin axis, on a true running wheel.
Sometimes a definition for a loaded tire is used. In that case the radial
run out is the variation of the loaded radius. In this measurement the
radial stiffness variation around the circumference plays a role. This
method is replaced in most cases by both the measurement of unloaded
radial run out and the measurements of force variations with a fixed axle
height.
The lateral run out is the variation in position measured parallel to the
spin axis at the point of maximum tire section width, on a true running
wheel.
These measurements can be complicated by the presence of trade marks
and labeling on the side walls.
Usually run out measurements are carried out at low speeds with me
chanical spring loaded feelers. For high speed measurements these are less
suitable on account of vibrations and an eventual loss of contact.
At the Vehicle Research Laboratory in Delft measurements are carried
out by contactless methods, fig. 8.7.72.
The radial run out as well as the growth of the outer diameter of cross-
ply and radial tires have been measured. From 10 to 180 km/h the in
MEASUREMENT OF TIRE PROPERTIES 615

crease in diameter was 1 to 3 mm for the radials and 10 to 12 mm for the


cross-ply tires.
The run out signal appeared to stay fairly constant and as it is periodic,
this signal can be represented by a series of sine waves each at a frequency
equal to an integer multiple of the wheel frequency.
Each sine wave component is called a harmonic of the wheel frequency
and when all harmonics are summed with their proper phase relation
ships, the original run out curve is reconstructed.
The amplitudes An of the harmonics are the magnitudes of the Fourier
coefficients of the periodic wave form expressed in the form:

F(t) Am sin (2-nn.f+

in which f = wheel frequency and <£„ = phase angle.


Analyzing the runout signals the following conclusions can be drawn:
—the fundamental (or first harmonic) is usually 3-5 times the ampli
tude of higher harmonics.
—a limitation of about 0.05 mm for the amplitude may be found as a
reproducible minimum value.
—with this limitation, usually 4-6 harmonics can be considered.
For example: 1st harmonic—A1 =- 0.2 -0.6 mm
2nd harmonic—A2 = 0.07-0.20 mm
3rd harmonic—A3 = 0.05-0.10 mm.
These results are from measurements of 2 radial and 2 cross-ply tires and
need not be generally valid [78].
B. Force variations
When a tire is loaded against a drum and held at a fixed axle height, the
vertical load will show variations for the free rolling condition. There also
appear to exist longitudinal and lateral force variations, although the cam
ber and slip angle are zero.
These force variations are due to the nonunifonnity of the tire.
Furthermore there can be moment variations, of which the self aligning
torque variation is the most commonly mentioned.
B-l. Radial force variation
A radial force variation can be the result of radial run out, variation in
stiffness around the circumference or a varying mass distribution, the lat
ter especially at higher speeds. These factors are partly interrelated.

3 rd harmonic
RF- radial fundamental
RC - radial composite

FIGURE 8.2.73. Typical graph of the radial force variation.


616 MECHANICS OF PNEUMATIC TIRES

lateral
tore.

angle of rotation

SFL . >IO> force level


LF - lateral fundamental
LC . lateral composite

FIGURE 8.2.74. Typical graph of the lateral force variation.

The force variations repeat themselves with each revolution of the tire
and can be considered as a periodic signal when the rolling speed is con
stant.
A typical graph of the radial force variation is shown in fig. 8.2.73.
The peak-to-peak value of the composite signal is a criterion for the
nonuniformity, but sometimes a maximum acceptable level for the ampli
tude of the first harmonic (fundamental) is indicated.
Generally the amplitude of the fundamental is smaller compared to the
peak-to-peak value. Theoretically this is not necessary.
The amplitude of the higher harmonics generally decreases with in
creasing order of harmonics.
Radial force variations are generally slightly speed dependent [79, 80] as
is shown in fig. 8.2.85. Essential for this type of measurement is a very ac
curate balancing of the tire and rim, as an unbalance of only a few grams
(0.01 N) will result in force variations at high speeds in the same order of
magnitude as the nonuniformity force variations to be measured.
Changing the load or inflation pressure will change the force variations
[81]. The ranking of different tires with respect to force variations can also
be changed.
Radial force variations acceptable for car manufacturers range from 100
to 150 N for the peak-to-peak value for the usual European tire sizes for
13" or 14" rim diameters.
B-2. Lateral force variation
For an ideal uniform tire the free rolling lateral force variation is zero.
Usually however there exists not only a variation in the lateral force, but
the average value differs from zero during one revolution.
In fig. 8.2.74 a typical curve is shown.
Lateral force variation is slightly speed dependent like the radial force
variation [79, 80], fig. 8.2.85.
The lateral force variation will almost invariably depend on the direc
tion of rotation, especially for radial-ply tires.

FIGURE 8.2.75. Tire approximation as a truncated cone.


MEASUREMENT OF TIRE PROPERTIES 617

LFV»

a . lateral force level for right rotation


b i lateral force level for left rotation
conicity c e
a <• b
ply steer h =

FIGURE 8.2.76. Side force variation depending on the direction of rotation.

The average value about which the force variation occurs, can be di
vided, according to the cause in:
—conicity
—ply steer.
Conicity
This term was apparently derived by considering a tire to assume the
shape of a truncated cone as illustrated in fig. 8.2.75. Based on geometry
such a configuration would generate a force towards the apex of the cone
regardless of which direction the tire or "cone" was rotated. Thus conicity
by definition is a force component which does not change direction with
reverse rotation when measuring tire lateral force variations.
Similar forces would be generated by reversing a tire having camber to
the road surface. Therefore conicity is often referred to as "pseudo cam
ber" [82].
Ply steer
This component of the lateral force variation describes the influence of
the plies in a tire in generating forces which could steer a vehicle from its
intended straight line course. Ply steer, by definition, is a force component
which changes direction with reverse rotation, when measuring the tire
lateral force variations.
Similar forces would be generated by a tire operating at a small slip
angle.

ply steer

FIGURE 8.2.77. Distribution of conicity and ply steer forces for a given tire construction.
618 MECHANICS OF PNEUMATIC TIRES

Ply steer forces are therefore often referred to as "pseudo slip". In fig.
8.2.76 a typical force variation curve is shown.
If a tire shows a conicity effect without ply steer (a = — b), then there
exists a constant lateral force, independent of the direction of rotation.
Turning the tire on the rim will cause a change in the direction of the side
force.
A tire that shows ply steer without conicity effect (a = b) produces side-
force in one specific direction that changes only by changing the rotation
of the tire from forward to backward. Turning on the rim or mounting at
the right or left side of the car does not change the direction of the side
force.
Conicity has the character of a random effect for a tire design, while the
distribution of ply steer forces for a given tire design is not random, fig.
8.2.77 [83].
Ply steer forces appear to be nearly identical in magnitude for a given
tire construction. Conicity seems to be due to mainly an incorrect position
ing of the belt, off-center belts, and partly to random irregularities [82].
A belt off-set toward the white side wall causes a conicity force towards
the white side wall. There also appears to be a related asymmetrical drag
distribution which results in an aligning torque, aggravating the nonuni-
fonnity.
Off-center belts result in a difference of circumference between shoul
ders which can be observed when the tire is dissected.
Ply steer of radial-ply tires depends predominantly on the construction
of the belt layers. A belt consists of 2 or more layers of rubberized parallel
cord materials, called plies.
The belt cords lie generally at an angle of 12-20 degrees to the crown
centerline.
A belt that is symmetrical with respect to the thickness midplane will
give low ply steer values. Examples of this are belts B and F of Fig. 8.2.78.
A belt very often has an alternating stacking of the cord layers with re
spect to the cord angle. With regard to the thickness midplane this is
asymmetrical and due to this fact ply steer forces will appear in a specific
direction depending upon the belt cord configuration. Examples of this are
belts A and E of Fig. 8.2.78.
This effect is due to different behavior of symmetrical and asymmetrical
laminates under tensile stresses which occur in the contact area of a
loaded tire [84].
The flattening of a bent belt with an asymmetrical stacking of cord lay
ers such as belts A and E, will result in a deformation in the contact zone.
A belt element deforms in such a way as to cause a small slip angle which
in turn is responsible for the ply steer forces. See Fig. 8.2.79a.
The presence of carcass layers with cord angles of 88-90 degrees can
cause small asymmetric effects. Normally the belt is built on the carcass
plies, as shown in section A of Fig. 8.2.78. To obtain a more symmetric
tire, one may locate the belt between the carcass plies as shown in sections
C and D of Fig. 8.2.78. This construction gives a slight improvement in
ply steer forces, as shown in the data for tires B and D of Fig. 8.2.80, taken
from [84]. One may conclude from this that the main cause for ply steer
forces is the construction of the belt.
The conicity and ply steer forces of the different constructions shown in
fig. 8.2.78 are given in fig. 8.2.80.
Another approach to achieve a radial tire without ply steer is through a
belt that has lateral sections with different ply steer responses that are self
cancelling. For example, in a split response belt construction (Fig. 8.2.81)
MEASUREMENT OF TIRE PROPERTIES 619

a) cord
angle
positive

midjalane

stiffness symmetry with regard to the


thickness midplane resulting in low
ply steer forces

asymmetrical design resulting in


relatively high ply steer forces

different carcass layer


d) configurations

C D A
FIGURE 8.2.78. Different stockings of bell layers.

the belt might be laid up so that the right half alone would cause a positive
ply steer force and the left half alone would cause a negative ply steer
force, with the result that for the whole belt a zero ply steer would be ob
tained.
Maximum acceptable levels for car manufacturers for the peak-to-peak
values of lateral force variation range from 70 to 150 N, for the usual sizes
165-13 and 175-14.
The maximum acceptable level for conicity forces range from 70 to 100
N. •
Influence of load, inflation and rim width
The influence of normal load at constant inflation pressure is given in
figure 8.2.82a for the ply steer forces and in figure 8.2.82b for the conicity
forces [82].
620 MECHANICS OF PNEUMATIC TIRES

M NO PIT SIEER RMNAl Tint

FIGURE 8.2.79. Distortion of tread region against the road.

The ply steer force depends on vertical load like a tire rotating at a slip
angle. Using data from measurements on a 33 inch diameter wheel, con-
icity and ply steer forces increase as the vertical load increases. At con
stant vertical load the conicity depends strongly on inflation pressure, Fig.
8.2.83a. A relatively small increase in inflation pressure may reduce the
conicity force remarkably. Ply steer forces are rather independent of infla
tion pressure, Fig. 8.2.83b, [82].
The influence of rim width on conicity and ply steer forces is small [82].

FIGURE 8.2.80. Conicity and ply steer forces of the different belt'constructions shown in
figure 8.2.78.
MEASUREMENT OF TIRE PROPERTIES 621

FIGURE 8.2.81. Split response belt construction resulting in no ply steer forces.

Effect of conicity and ply steer on vehicle performance


The effect of ply steer and conicity can be judged by considering the
"pull" and "drift" behavior of the vehicle. Vehicle jmU refers to a condi
tion where the driver must apply a constant torque totHe steering wheel in
order to maintain a straight-line course. If the steering wheel is released
the vehicle gradually changes course with the steering wheel returning to
the centered position. Pull can be judged unacceptable if, for instance, on
a level, divided highway the vehicle would change lanes in 1/10 mile or
less when the steering wheel is released, at a speed of 80 km/h.
Vehicle drift can be defined as a vehicle operating at a constant yaw or
side slip angle to the direction of motion, the rear wheels do not follow in
the exact path of the front wheels [82].
Effect of conicity
The direction of conicity force cannot be predicted. If the net conicity
force on both front wheels is not zero, this will result in a pull of the ve
hicle. The driver has to correct with the steering wheel and this can be
fatiguing to the driver. If the total conicity force on the front axle is kept
below 180 N then most vehicles will show satisfactory behavior [82].
A conicity force on the rear axle will cause a slight off-set of the rear
wheels, they will not follow the same path as the front wheels. The driver
will note a change in the steering wheel centering, but no vehicle pull will
be noticed, assuming that there are no resultant conicity forces on the
front axle, or the conicity forces are equal and in opposite directions [82].
In summary, the conicity forces on the front axle can give rise to unde
sirable pull problems of the vehicle which may be very irritating to the
driver.
622 MECHANICS OF PNEUMATIC TIRES

HR78-15 RADIAL
CONSTANT INFL-28 PSL

INORMAL LOAD-IBS. |
Fig. 8.2.82a

HR78-15 RADIAL
CONSTAN1 INFL-28 PSL

">
s

TOOd
| NORMAL LOAD-LBS.

Fig. 8.2.82b
FIGURE 8.2.82 Influence of normal load on ply steer and conicity for HR78-15 tire at con
stant inflation pressure.
MEASUREMENT OF TIRE PROPERTIES 623

I'll.

HR78-15 RADIAL
CONSTANT LOAD-
•1280 LBS.

I INFLATION- PS I

Fig. 8.2.83a

70- TIRE E
TIRE H

HR78-15 RADIAL
CONSTANT LOAD -1280

| INFLATION- PSI

Fig. 8.2.83b
FIGURE 8.2.83 Influence of inflation pressure on ply steer and eonicity at constant load for
HR78-15 tire.
624 MECHANICS OF PNEUMATIC TIRES

Effect ofply steer


Ply steer forces are nearly identical in magnitude for a given tire con
struction. A vehicle equipped with four tires of the same construction will
show a slight drift due to the ply steer forces. All tires will show a ply steer
force of the same magnitude and in the same direction. No resultant pull
will occur in this case and the driver may hardly notice the small yaw
angle of the vehicle, Fig. 8.2.84a, [82].
Extreme test conditions can be achieved by building a special tire with a
reverse tread ply construction, resulting in ply steer forces in the opposite
direction.
Two of these tires mounted on a rear or front axle with the original pro
duction tires on the other axle will result in the same drift of the vehicle
but there will be a considerable steering wheel off-set from the normally
centered position, Fig. 8.2.84b. No noticeable pull will occur.
Thus in summary it appears as if the ply steer forces for a given con
struction manifest themselves as a slight drifting of the vehicle, as in
dicated in fig. 8.2.84 and that mixing tires of different tread ply construe-

DIRECTION OF TRAVEL

I I

FIGURE 8.2.84 Effects ofply steer (a) and reverse tread ply on the front tires
(b) on a vehicle.
MEASUREMENT OF TIRE PROPERTIES 625

DIRECTION OF TRAVEL VEHICLE


\ ATTITUDE

FIGURE 8.2.84 (con't)


tion on front and rear axles may affect the steering wheel centering. Ply
steer forces are not noticed in normal applications.
B-3. Longitudinal force variations
In the longitudinal direction a tire will produce nonuniformity force
variations as well. The variations will be around an average value corre
sponding to the rolling resistance force. The longitudinal force variations
are partly related to the radial run out and to the radial force variations.
Of the three force variations only the longitudinal force variation
strongly increases with speed. At low speed it tends to lag behind the
radial force variation by 90 dog.; the lag increases as speed rises [82].
The usual tire uniformity machines are able to measure radial and lat
eral force variations at low speed (60 rpm). The measurement of longitudi
nal force variation on low speed machines is less meaningful, considering
the strong sensitivity to speed [79, 80, 85] Fig. 8.2.85. In ref. [86] an inves
tigation into the causes of steering wheel vibrations is described. These vi
brations are strongly correlated with the longitudinal force variations
(70%) and for a minor part (30%) to radial force variations. These relations
626 MECHANICS OF PNEUMATIC TIRES

PEAK/PEAK
FORCE LB
100
RADIAL
10

LONGITUDINAL

LATERAL

50 100
SPEED MPH
FIGURE 8.2.85. Ranges offorce variations depending on speed.

were obtained by means of statistical handling of the measured data.


At present there are no accepted standards for either longitudinal force
variation levels, or the variation in the tire moments such as aligning
torque.

Separating tire and wheel force variations


For several reasons it may be desirable to separate the force variations
induced by the wheel and the tire. The results may be used in order to
match the tire and wheel and thereby minimize the assembly nonuni-
formity. The knowledge of such individual force variations is particularly
useful when no true rims are available for tire measurements.
A method of separating the first harmonics will be described [87].
Consider a tire wheel assembly which has a first harmonic radial force
variation of 120 N and an angular location of 80 deg. from the valve stem.
After this measurement is taken the tire is rotated by 180 deg. on the rim
and measured again. Now the result may be 170 N at 155 deg.
The first harmonic of the tire and the rim and their position can be de
termined as indicated in fig. 8.2.86a.
The vector representing the original condition of the assembly has a rel
ative length of 1 20 N at an angle of 80 deg. The "after rotation" condition
is represented by the other vector.
The next step is to connect the two vectors with a straight line. A vector
drawn from the origin to the midpoint of the line represents the contribu-
MEASUREMENT OF TIRE PROPERTIES 627

r not AIICN of wm mm

UIUUI1
(OtIGIKAl CONDITKDNI
FOKI nil II LI M TIM. It II «T If
H- lion VI . I }TU IOIIGINAL POSITION ON WHEtl)

ASHMIIT \ /TiH nu Aim*


(Aiiei Tin «n«i . V (AITII IH«0 IOTAT1D lir ON mill)
(OKI MA« 1) II AT ISS

TTBIMIAIIOS'N
AIIfMIlT
(AfTEl MATCHING Tllf TO WNHU.
-

FIGURE 8.2.86. Separation of tire and wheel force variation.

tion of the wheel to the first harmonic amplitude and angular location, in
this case 120 N and 125 deg. respectively.
The vector from the midpoint up to the original vector represents the
contribution of the tire to the first harmonic amplitude and its angular po
sition in the original condition, which scales at 90 N and 15 deg. respec
tively, Fig. 8.2.86b.
After these measurements it is easy to position the tire on the wheel in
such a way that the first harmonics of tire and wheel are in opposite direc
tions. In the case considered a remarkable improvement could be
achieved. The assembly had a force variation of only 2 N, Fig. 8.2.86c.

C. Unbalance
A tire usually shows a certain amount of unbalance. Theoretically this
does not present any great problem as the tire can be balanced by lead
weights. If this is properly executed, statically and dynamically, the tire
behavior is very uniform, assumed that there are no dimensional nonuni-
formities or force variations.
The presence of a large amount of unbalance, however, usually in
dicates that other nonuniformities exist. As a rough indication for the uni
formity of a tire the amount of unbalance may be considered and can be
measured very easily.
In some regulations maximum unbalance levels are indicated. DIN
7817 requires, for example, an unbalance of less than 700 gr. cm. For a 14
inch rim this means a maximum unbalance weight of about 45 gr.
Unbalance vibrations can be caused either by tire unbalance or by un
balance of vehicle components.
To separate both effects a similar method can be followed as described
earlier, viz. by separating first harmonics of wheel and tire.
In this case a wheel-tire combination is rotated on the hub by about 2 or
3 bolt holes. It is not necessary to rotate it by 180 deg. In fig. 8.2.87 this
method is shown [88].
628 MECHANICS OF PNEUMATIC TIRES

IS! BOLT

INITIAL POSITION is.' BOLT


(is! BOLT AND AFTER 144° CLOCKWISE
BOLT HOLE COINCIDE) ROTATION OF WHEEL
fU (2)

VEHICLE
COMPONENT

is! BOLT HOLE

TIRE
COMPONENT

VECTORS REFERENCED TO SEPARATION OF TIRE 4 VEHICLE


121 BOLT HOLE COMPONENTS
(3)
FIGURE 8.2.87. Separation of tire and vehicle components unbalance.

Wheel Uniformity
Car wheels are subject to manufacturing fluctuations and contain some
degree of nonuniformity. The maximum radial and lateral runout at the
beads should not exceed approximately 1.25 mm. (DIN 7817). Accurate
wheel centering contributes greatly to a good degree of uniformity.
Tire uniformity measurements are made either on precision rims or car
wheels.
Figure 8.2.88 shows car wheel radial and lateral run outs as measured
on the beads, together with the corresponding free tire run out as mea
MEASUREMENT OF TIRE PROPERTIES 629

sured on the side walls of a preconditioned tire. Lateral run out spectra of
wheel and tire sometimes show a good correlation.
Besides the lateral and radial run out it is essential that the bolt holes be
concentric with respect to the rim. In production this requirement cannot
generally be fullfillcd with a high degree of precision.
Another problem is the concentric mounting of a wheel on to the hub
and its reproducibility. In [88] a check on a Chevrolet revealed that for
that car, the run out of the rims for different positions on the hub was
within 0.5 mm.
An eccentric mounting of a wheel-tire assembly with a mass of 10
kg and eccentricity of 0.25 mm causes an unbalance force variation with
an amplitude dependent upon the speed of the vehicle. At 15 revolutions
per second (which corresponds to about 90 km/h) this unbalance is:
m • u2 • e - 10 • (2W-15)2 • (0.25)10-3 = 22 N
For a rim radius of 0.15 m this corresponds to about 10 grams in terms of
unbalance weights.

Effect of Tire Nonuniformities on Vehicles


The tire-excited vehicle vibrations introduced by the nonuniform prop
erties of the tire are felt as vibrations of those parts of the car which are in
contact with the passenger, and heard as sound in the passenger com
partment. It appears that no car is insensitive to tire force variations but
some are more sensitive than others. The vibrations of vehicle systems that
may be influenced by tire nonuniformities are listed on the following page.
(SAE definitions, [cf. ref. [3] sec. 8.3])
shake —a mode of vibration of the entire sprung mass as a flexible
body occurring generally in the frequency range of 5-25 Hz.
Speed range 80-130 km/h.
flutter —forced oscillation of steerable wheels about their steering
axes, frequency range 8-20' Hz, speeds 80-140 km/h.
wobble — a self-excited oscillation of steerable wheels about their steer
ing axes occurring without appreciable tramp. Frequency
range 4-8 Hz, speeds 40-100 km/h.
shimmy — a self-excited oscillation of a pair of steerable wheels about

Tirt

Whtel I
Petition
579
3579 11 1 3579
3 5 7 9 II 1
( W.S.W)
Lateral Runout I BS.W )

fir.

liformityH
machine drum Wh'ctl >
Position
1 3579 II 13579 II I

Radial Runout
FIGURE 8.2.88. Radial and lateral run out measurements of wheel and tire.
630 MECHANICS OF PNEUMATIC TIRES

their steering axes, accompanied by appreciable tramp. Fre


quency range 4-8 Hz, speeds 40-100 km/h.
thump — a periodic vibration and/or audible sound generated by the
tire and producing a pounding sensation which is synchronus
with wheel rotation. Frequency 30-60 Hz, speeds 30-110
km/h.
roughness —vibration perceived tactily and/or audibly, generated by a
rolling tire on a smooth road surface and producing the sen
sation of driving on a coarse or irregular surface: Frequency
30- 1 30 Hz, speeds 50- 1 1 0 km/h.
harshness —vibrations perceived tactily and/or audibly, produced by in
teraction of the tire with road irregularities. Frequency 15-
100 Hz.
hop —the vertical oscillatory motion of a wheel between the road
surface and the sprung mass.
tramp —the form of wheel hop in which a pair of wheels hop in oppo
site phase.
Another type of vibration is waddle, which is combination of nosing
and rolling motion of the vehicle. This motion is very sensitive to tire non-
uniformities. Because the lateral and radial forces vary with each wheel
revolution, and the effective rolling radii of left- and right-hand wheels are
not exactly the same, the resulting motions may be in phase or in opposite
phase. The effect may be felt through the steering wheel and experienced
on flat roads as low frequency oscillatory forward motion combined with
slight rolling motion. This low frequency nosing action of the vehicle can,
for instance, be recorded with a drift angle meter attached to a front
wheel, as described in section 8.3.3, figure 8.3.43.
In fig. 8.2.89 the relation between frequency, vehicle speed and nonuni-
formity harmonics is given theoretically. As a norm a tire size 175-14 with
a radius of R = 300 mm has been taken and the range for a small tire (R =
250 mm) and a big one (R = 350 mm) is indicated for the first harmonic.
' In order to find the relation between the harmonics and certain vibra
tions it is necessary to know the frequency of the vibrations along with the
vehicle speed and the effective rolling radius of the tires.

FIGURE 8.2.89. Relation between frequency, vehicle speed and nonuniformity harmonics.
MEASUREMENT OF TIRE PROPERTIES 631

For low frequency vibrations, below 20 Hz, that occur at speeds of 80-
120 km/h the first harmonic of the force variations is the important quan
tity.
At relatively low speeds of about 50 km/h the 4th and higher order har
monics reach frequencies of over 35 Hz.
This means that the "tire generated" frequencies are for a substantial
part filtered by the wheel suspension system, depending upon the natural
frequencies and damping ratios in the different parts of the wheel suspen
sion and steering system. Resonant frequencies in the spindle axle—steer
ing wheel system, for example, are in the range of 25 to 35 Hz for a series
of 1973 models of passenger cars [86].
On a car driving on a straight road the vertical force variations cause a
variation in axle height. At a constant speed the angular velocity of the
front wheels must be constant. The axle height variations will cause angu
lar decelerations and accelerations in order to keep the vehicle speed con
stant.
Tire Uniformity Grading Machines
Numerous tire uniformity machines are described by various authors
and many machines are in use. Figure 8.2.90 is a schematic of a tire uni
formity machine [89]. The test tire is mounted on a precision rim with pro
vision for rapid tire inflation. The tire is loaded against a drum of the larg
est size that is practical, since the roll size affects the force variations [90].
The tire is run at zero camber and zero slip angle at speeds less than 60
rpm. This low speed of 60 rpm or 1 rps has been chosen so that the ampli
tude of the tenth harmonic of the force variations can be reliably mea
sured. This requires the minimum machine and tire resonant frequencies
to be at least four times the frequency of the tenth harmonic, or 40 times
the rotational frequency. Since the lowest tire resonance occurs at approxi-

souo OB SPLIT m

not LOU
FIGURE 8.2.90. Tire uniformity machine schematic.
632 MECHANICS OF PNEUMATIC TIRES

mately 40 c.p.s., it is necessary to restrict the maximum wheel speed to 1


rps [90, 91].
Reproducible results for both forward and reverse rotations are only ob
tained when the tires are properly conditioned. The resulting force varia
tions are recorded at the specified inflation and load, with a fixed distance
between the tire and drum axle. This constant height method is employed
since a constant load will cause axle height variations for nonuniform tires
with resulting inertial and factional errors in load application.
Some machines can measure radial, lateral and tangential force varia
tions in tires, as well as free radial runout variations, free width variations,
and the crown thickness. Free radial runout variations are measured either
on the centerline or on both shoulder ribs. The free measurements can be
simultaneously executed when force variations are measured, the dis
placement transducers being located a sufficient angular distance from the
footprint area to avoid distortions from loading [89]. However, the mea
surement of free radial runout is sometimes conducted on special ma
chines with a preconditioned unloaded tire. A displacement transducer
fixed in space contacts the crown of the slowly rotating tire.
To measure higher harmonics of tire nonuniformities the test rig must
be designed for high natural frequencies in the directions in which the
measurements will be carried out. This requirement implies a light weight,
very stiff test rig. A measuring system based on foil strain gages has the
disadvantage that relatively high deformations are necessary. Semi
conducting strain gages offer an improvement [79] and force transducers
consisting of piezo-electric measuring cells are even better. These types are
sometimes used in laboratory measurements.
The mass of the structure between the contact area and the measuring
system should be as small as possible. For example, take the mass of the
tire-wheel assembly, axle and bearings as 50 kg. The measuring system is
mounted outside the bearings. At 20 Hz (120 km/h) a movement of 0.05
mm of the axle appears to occur. In this case the inertia forces introduced
in the measuring system are:
m • w2 • z = 50 • (27r20)2 • (0.05) • 10'5 = 40 N
With such equipment it is hardly possible to measure the first harmonic of
the force variations which are small compared to the 40 N due to inertia
forces.
In Europe the requirements for radial and lateral runout, and force var
iations are standardized, for instance, in the W.d.K. instructions (Wirts-
chaftsverband der deutschen Kautschukindustrie) [92], which are very sim
ilar to the SAE recommendations. These standards are often required by
the car manufacturers. The maximum level for the radial and lateral run
out is 1 .5 mm while in the DIN standards a maximum radial and lateral
runout for rims is indicated to be 1.25 mm (DIN 7817). Usually the car
manufacturers have their own requirements that range for radial runout
from 1 to 1.5 mm and for lateral runout from 1.2 to 1.5 mm.
In the W.d.K. instructions the conditions for the force variation mea
surements are given. The diameter of the drum should be 854 mm like the
SAE recommendations and the prescribed speed is 60 rpm for car tires
and 20 rpm for truck tires. This corresponds to road speeds of 10 and 3
km/h respectively. The inflation pressure for car tires is 2.0 bar and the
recommended loads depend on the tire size.
MEASUREMENT OF TIRE PROPERTIES 633

Corrections for tire nonuniformity


If a tire shows an unacceptable level of nonuniformity a few measures
can be taken in order to minimize the effects. All measures are restricted to
radial runout and radial force variations. Lateral force variations are diffi
cult to remedy.
Matching
In general both tire and rim have nonuniformities. It is possible to
match the tire and the rim in such a way that the nonuniformity of the as
sembly is minimized. The force variations can be divided into tire and
wheel components and the first harmonics can be placed 180 deg. apart.
The results obtained in this way have proved to be very encouraging [87,
88].
For normal production, uniformity machines are used that indicate the
lowest point of a wheel, averaged over both bead seats and the high point
of a tire mounted on a true wheel. To match them the tire is positioned on
the wheel such that the two points are adjacent. An example of the results
that can be achieved is given in figure 8.2.91 [87].
Grinding
Another way to correct tire uniformity is to remove rubber from the
tread. The radial runout can be improved by grinding the tread (truing)
with a fixed position of the grinder tool with respect to the spin axis. In
this treatment small amounts of rubber are removed at the high points of
the tread. The result is a true tire without dimensional runout. In cases
where an extreme radial runout occurs it is desirable to reduce the runout
to an acceptable level in order to prevent an extreme initial "wear" of the
tread pattern.
The force variations become smaller but usually do not disappear en
tirely. The next step could be the removal of tread rubber at places where
the highest force variations occur. In this way a radial runout is in
troduced in order to decrease the radial force variations. The removal of
rubber can be made completely across the tread width, but the major por
tion of the effect of a radius reduction occurs at the shoulder area of the

PERCENT

FIRST HARMONIC OVER 20 LI FIRST HARMONIC OVER 30 IB

FIGURE 8.2.91. Results of matching tires and rims.


634 MECHANICS OF PNEUMATIC TIRES

-.010 INCH
INITIAL

SHOULDER
FREE RUNOUT CONCENTRIC
CORRECTED

RADIAL FORCE
VARIATION

•10 LBS

/A)k / CONCENTRIC
t 1 INITIAL
FIRST HARMONIC
RADIAL FORCE \
VARIATION ^ / ^ S js CORRtCltU
()NE TIRE
i[EVOLUTION
FIGURE 8.2.92. Uniformity recording, two step correction.

tread. This is due to the relatively high radial stiffness of the shoulder as
compared to the tread center [81]. Results of this two-step improvement of
uniformity are shown in figure 8.2.92. Truing is one of the simplest meth
ods to improve the tire uniformity. It will result in a reduction of vehicle
shake, although not necessarily to an acceptable level. As a rule no more
than 0.5 mm will be removed.
A simple method for further improvement is eccentric machining. In
this case the first harmonic of the radial force variation is detected and the
movement of the grinding or cutting tool is so adjusted that an eccentric
run out of the tire circumference is obtained. Fig. 8.2.93. The highest point
of the tire is at the lowest point of the first harmonic. This method is par
ticularly suited for service conditions. A more complicated method is the
application of servo controlled positioning of the cutting tool. This method
allows considerable flexibility of the control signal used to modify the free
runout. The position of the cutting tool can follow the force variations
very closely. Usually the corrections are made in the shoulder area of the
tread. Restrictions are applied to prevent an overfeeding of the grinders.
On-car grinders have been developed that are based upon a similar
principle [93]. A comparison of trued tires and force corrected tires is
given. In this test the results are compared over 70% of the tread life on a
GM proving ground. It appeared that trued tires given an improvement
for the first 20,000 miles. This corresponds with about a quarter of the
tread life. After this period the force variations were at the level at which
the original tires were before truing. The force-corrected tires proved to
have a permanently lower force variation level.
Vehicle shake, which is strongly related to the first harmonic of radial
MEASUREMENT OF TIRE PROPERTIES 635

ECCENTRIC CAM
ADJUSTABLE PHASE AM AMPLITUDE

I? I'.CH DIAMETER ROAD DRUM

CUTTING TOW

RUMDUI TRANSDUCER

CUTTING TOOt PIVOT

f—LOKD Aluf PIVOT

FIGURE 8.2.93. Eccentric correction machine schematic.

force variation, is much reduced after the correction of the tires, fig.
8.2.94a. Roughness, which is related to longitudinal force variation, is not
necessarily reduced and in some cases roughness may even become worse,
fig. 8.2.94b [81]. The measurements in these tests are carried out with ex
perienced drivers on a judgement basis. With specially trained and se
lected drivers these subjective ratings can be very reliable and reproduce-
able [78]. A subjective rating scale is established in SAE Recommended
Practice, SAE J1060, [94].
All the tire correction methods described above have the disadvantage
that the tread depth of the tires is reduced, thereby decreasing the tread
life. For steel belted tires this reduction of tread life seems to be accept
able, provided that the amount of rubber removed is limited. Another dis-
100

10 20 30 40 SO 60
RADIAL FORCE VARIATION - POUNDS

FIGURE 8.2.94a. Improvements of radial force variation by servo-controlled correction.


636 MECHANICS OF PNEUMATIC TIRES

£
o
ae

12 SETS
TIRE-WHEEL
ASSEMBLIES

ACCEPTABLE UNACCEPTABLE
VEHICLE RATING (10 - 1 SCALE)

FIGURE 8.2.94b. Change of vehicle ratings by servo-controlled correction.

advantage is that the costs to correct the tires in an assembly line at the
automobile plant are considerable.
8.2.7 Tire Flaws and Separations
by Stephen Bobo
Introduction
The optimum tire is one of uniform rotational symmetry having no
characteristics which will cause vibration, shimmy, or other running per
turbations. However, by the nature of their construction, tires have anom
alies caused by splices, belt misplacement, uneven cords, cord angle and
cord count variation and other conditions. It is therefore necessary to keep
these potentially harmful variations within controlled limits during the
manufacturing process, but certainly before the tires enter service. A po
tentially valuable method for finding anomalous characteristics within the
tires is nondestructive inspection.
Until recently, inspection of tires for flaws and separations has been a
visual and tactile operation. Experienced tire inspectors have achieved re
markable skill in identifying irregularities within the tire. Tactile in
spections on the inner surface of a new tire can reveal bead defects, stray
cords, lapped plies and separations, while on older tires it is possible to see
evidence of separations and belt damage.
However, with the advent of low cost, nondestructive inspection meth
ods, together with the need for safer, longer lasting tires, some of the meth
ods used in other industries for insuring product quality have been
adapted to tire production. [95].
An ability to find flaws in tires implies the need to relate these to tire
MEASUREMENT OF TIRE PROPERTIES 637

performance. The primary source of flaw-failure correlation in the indus


try comes from returns from consumers for adjustments. These tires are
analyzed for cause of failure by skilled inspectors generally situated at ma
jor distribution centers, with reports being sent to company headquarters
for tabulation.
In order to be able to predict service performance of new designs, tire
manufacturers have historically relied on test vehicles, trailers and road-
wheels. One source has stated that to infer the performance of tires from
such tests, it is necessary to run a statistically valid population of tires as
many as 10,000 miles. The costs associated with such tests have limited
their use to relatively small numbers of tires.
In fact, the only moderately practical test of reasonable populations of
tires seemed to be by using roadwheels. A variety of test wheels have been
developed, the best of which were described earlier. These devices are of
value to the industry since tire performance can be characterized at far
less expense than would be possible if the data were to be gathered by
road experience. Their limitation stems from the fact that there is not
adequate statistical data relating the tire failures on a drum or flat belt ma
chine to road failures. Owing to the fact that there is not a steady flow of
air around the road wheel, as in road tests, the tires on roadwheels tend to
become warmer and therefore roadwheel tests are generally considered
thermal tests.
The roadwheel test cannot be considered nondestructive, nor is it en
tirely devoid of expense. Its use therefore is not entirely feasible for large-
scale testing of new tires to insure conformance to specifications.
For this reason testing methodology consisting of employing non
destructive inspection techniques correlated against roadwheel tests has
been established. The inspection techniques used have tended to fall into
three categories, optical, penetrating radiation, and ultrasonic holography.
The predominant optical technique in widespread use now is holog
raphy. A concomitant to the invention of the laser, holography (Holos:
greek for entire or whole) has become a tool for identifying dimensional
changes in the order of wavelengths of light [96].
Use of holography in tires has been primarily associated with locating
separations and areas of unusual stress within the tire, although some
work has been done in visualization of tire vibrations and resonances, as
for example, the work of Potts [97].
Most tire holography has been done with equipment developed by
Brown and Grant under patents now owned by Newport Research Corpo
ration, Fountain Valley, California.
Approximately 50 tire holographic machines are now in use throughout
the tire industry and the feasibility of their application in nondestructive
tire inspection has been established [98].
Figure 8.2.95 is a photograph of a tire analyzer. The instrument pro
duces double-exposure holographic interferograms in conjunction with
the vacuum differential straining technique. The tire analyzer holographs
the inner surface of the tires and can be operated in a completely auto
matic mode. The basic principle of the technique is that two exposures are
made of each quadrant of a tire, one exposure at atmospheric pressure and
one exposure in a partial vacuum. If a separation exists in the quadrant,
the air trapped in the separation will force the adjacent surfaces to bulge
out. In the reconstructed image of this hologram, fringes are generated hi
the area of the separation. The fringes are caused by an interference be
tween the two holograms generated from each exposure on the same pho
tographic film
638 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.95 Tire holographic analyzer

Figure 8.2.96 is a simplified schematic of the optical system used in the


tire analyzer. The holographic light source is usually a Helium-Neon Laser
operating in a single mode. The optical shutter located adjacent to the la
ser is electronically controlled by a photodiode in the film plane and as
sures the proper exposures. Reflectors direct the laser beam from the laser
through a window in the dome flange into the vacuum chamber. Here the
beam is split into the reference and object beams. Lenses are used to ex
pand the beams so as to illuminate the inner surface of the tire and the
photographic film. The film is handled and controlled automatically by a
film cassette which holds enough film to inspect about 15 tires.
Tires are prepared for inspection by first inserting four tire spreaders
across the tire beads at 90° intervals around the tire. The spreaders force
the beads apart and make the inner tire surface more accessible for ho
lographic inspection. Tires are allowed to relax in the spread position for
at least 20 minutes before they are loaded into the tire analyzer. This re
laxation allows creep of the rubber to subside. The tire is then placed on a
rotating table. The entire system is supported pneumatically and is iso
lated from any environmental disturbances. Once the tire is securely in
place, the automatic inspection cycle is started. The cycle begins with the
lowering of the dome to form a sealed chamber. The dome rests on a
flange and forms a vacuum tight seal. The film is advanced and one ex
posure is taken at atmospheric pressure. A partial vacuum is formed in the
chamber and a second exposure is made of the same quadrant of the tire.
The film is then advanced and the tire is rotated 90° so that the next quad
MEASUREMENT OF TIRE PROPERTIES 639

rant is in position to be holographed. This procedure is repeated until the


four quadrants have been inspected. The cycle is then terminated with the
lifting of the dome. Several seconds elapse between each function to allow
vibrations to dampen. The inspection for one tire requires approximately
2-1/2 minutes.
The photographic film is processed using standard developing chem
icals with no special care necessary. Holograms are reconstructed and ana
lyzed using an auxiliary reconstruction setup, with another helium neon
laser as the light source. The optical magnification and geometry is the
same as that used by the Tire Analyzer. This is necessary so that original
sizes are maintained and aberrations minimized. Holograms are read by
an operator and defect maps of the tire inner surface are constructed.
Figures 8.2.97a and 8.2.97b are typical interferograms generated by the
tire analyzer.
X-Ray
X-ray inspection has been employed in the analysis of defects in tires
for over 30 years. X-ray radiography was first used during world war II in
the inspection of aircraft tires in an attempt to determine the cause of tire
failure during take off. In 1958, X-ray fluoroscopy was applied to the ex
amination of tires and this technique has grown over the years until today
there are more than 60 units in operation throughout the tire industry,
monitoring component placement and builders defects or flaws in the
manufacturing process.
In X-ray radiography, the X-rays transmitted through an object pro
duce a photographic record on film. When a beam of X-rays strike an ob
ject, some of the X-radiation is absorbed by the object while the amount
transmitted is a function of the nature of the material, chemical composi
tion and thickness.
Fluoroscopy differs from radiography in that the X-ray image is viewed
directly on a fluorescent screen which can be enhanced with image in-

optical Shutter

vN 100%
Reflector

Pneumatic Tire
Dame Flange

FIGURE 8.2.% Schematic of Optical System of GCO holographic tire analyzer.


640 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.97 Typical holograms showing flaw location.


MEASUREMENT OF TIRE PROPERTIES 641
tensifiers or combined with television cameras which can display the im
age on a remote television monitor.
The X-ray units usually employed in the tire industry are considered
low power X-ray generators, generally operating in the range of 30 to 150
KV.
X-rays are produced when an electrically charged particle of sufficient
kinetic energy is rapidly decelerated. Conventional X-ray tubes produce
radiation by high speed electrons impinging on a metal target. The X-rays
emitted from the target consist of a mixture of wavelengths and the varia
tion of intensity with wavelength is a function of tube voltage, as seen in
Figure 8.2.98 for a tungsten target at several different voltages.
The intensity is zero up to the short-wavelength limit, which is given as
λSWL. ** 12.4/KV. As the tube voltage is increased, the intensity of all
wavelengths increases rapidly to a maximum and then decreases with no
sharp limit on the long wave-length side, while both the short wavelength
limit and the position of the maximum shifts to lower wavelengths, which
results in greater energy and more penetrating power. The intensity of the
continuous spectrum, which is used for radiography as contrasted to the
characteristic spectrum associated with X-ray diffraction, is lc,,NT =
AiZVm where A is a proportionality constant, i is the tube current, Z the
atomic number of the target material, V the voltage, and m is a constant
with a value of about two. Consequently, where larger amounts of white
radiation are desired, a heavy metal such as tungsten (Z = 74) is conven-

14

12

to

z 6

0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0


WAVELENGTH (X)

FIGURE 8.2.98 Intensity distribution with wavelength in the continuous X-ray spectrum of
tungsten at several voltages.
642 MECHANICS OF PNEUMATIC TIRES

tionally used as the target material. Increasing the tube current increases
the intensity of each wavelength for a given voltage but does not effect the
λSWL, and hence there is no change in the penetrating power. X-rays strik
ing matter are partly transmitted and partially absorbed. The intensity of
the transmitted beam Ix is given by the relationship Ix = /Oe~" where /<, =
the intensity of the incident radiation, x = the thickness of the material
and u is the linear absorption coefficient which is proportional to the den
sity ρ. The quantity u/ρ is a constant of the material independent of its
physical state, and is denned as the mass absorption coefficient, which is
the value usually tabulated for X-rays. Thus Ix = ItfT^"**. This ex
pression, while very exact for monochromatic and narrow (λ) beam radia
tion, holds only in a general way for the broad band of wavelengths used
in radiography or fluorpscopy where the absorption of a specimen de
pends not only on its thickness and density but to a very great extent on
the atomic nature of the material. Since short wavelength X-rays are of
primary consideration in radiography, the relationship

shows why this is such an important factor in radiography. A tabulation of


approximate radiographic equivalence factors, based on a compilation of
data from many sources is given in Table 8.2. 1 [99] and points out how
thickness, density, and the atomic nature of an object affects the radio-
graphic absorption. Two significant factors can be noted from the table:
relative absorptions of different materials are not constant, and the ab
sorption of a material is less dependent on its composition with increasing
voltage. Table 8.2.2 indicates the relative thickness equivalents at lower
voltages for rubber, aluminum, and steel and further illustrates these
points.
In tire applications, X-rays are usually permitted to diverge from their
source to some extent, thereby permitting wider coverage while still ensur
ing sufficient energy for penetration. The sensing surface, the film or phos
phor, is placed close to the tire and the internal structure of the tire is re
corded on the film or fluorescent screen.
Clarity of the picture is affected by a number of variables. In order to
permit the maximum definition, the following criteria should be applied:
1. The focal spot or source of X-rays should be as small as possible.
Generally, this is a function of machine design. However, a small
aperture source is desirable over a large aperture source, all other
considerations being equal.
2. The source of X-rays should be as far from the tire as possible.
3. The recording surface should be as close to the tire as possible.
4. The X-rays should be directed perpendicularly toward the recording
surface.
5. The area on the tire being inspected and the plane of the detector
should be as close to parallel as possible.
Other factors affecting clarity of image are related to the intensity and
wavelength of the source radiation, as previously discussed.
X-ray resolution is dependent on several physical variables. Geometric
unsharpness minimized by the rules set forth above is given by the rela
tionship
MEASUREMENT OF TIRE PROPERTIES 643

iimipi!};
f p p _ p -- **! p
01 \1 •| I-s
i •*
S
**M

•w
! "H,
V S- ~°
r~-
— p — 1
I
oc,
09 °D <*! g p p en
M «n • 8
« o
-3
a i 3
5 c 8. q
O
tflO d o p p ^ p ^ 2 <N so
I S 0

^ 1
cf.l Jl m rn p p — — — rn p so 1i -1§ 1
'3
o ••
1
§
Tt p p f) CS ?i rn p ev •s 1 i
m T3
i
CO
rt

*3
j
"S
m • u "~

1
c
AN oooz p p (N rn •n
Ij i.
'"3
J5
o

ri
i | |

! AN oooi p p (N rn p p
\ i.
ii ST H
•§
i
ti
.S
-Si
1
•5 AN (K)t' p p t «*J
*
m.
i• «> •S
id
^
o
^^
^
H ^

*
*S ^ d
X ANOJJ 3 00
R® S p p ^ n *
n p £ 1 a- 1
—— ® ® ® "* ^ —
r't
OO
<N
a* u ^ S ***
r-

m AN osi 2O
fN
so $o p o SO Tf V T n $ 1 al 8
3 ^ K
s |S|
]
fj
^j M o **- ?3
SO so
AN 001 O
p
rj es 00
£ tills
1
•t

j
t/
||| |£
so O CN
AN os O 1
|l||.a

I
8u ^o .2M xw "B"^
|

(2024
alumini- alloy
num)
Aluminum t
i 'J
mis^« Example:To"2!
ultiplied of
thicknes
tt Therefore, lea
giv •Torin
Magnesium Aluminum
Titanium X
Inconcl Zirconium
Uranium
I |
II
oo
1
6
C
N
*8
S
BO 3
644 MECHANICS OF PNEUMATIC TIRES

TABLE 8.2.2 Relative thickness equivalent


Voltage Rubber Aluminum Steel
25 KV 0.302 0.038 0.005
30 KV 0.995 0.114 0.014

where Up F, t, and D0 are defined and shown in Fig. 8.2.99. The geometric
unsharpness, which is the width of the fuzzy boundaries of shadows, can
strongly affect the appearance of the image.
Increases in tube current will increase the intensity and therefore
brighten a phosphor to a point, but not increase the penetration. Increas
ing the voltage will increase both penetration and brightness.
When it is necessary to resolve low contrast objects in the field, the abil
ity to see these defects depends on intensity and wave length of X-ray en
ergy as well as geometric considerations.
Obviously, in X-radiation diverging from the source, the farther away
the phosphor, the larger the image size. There is an optimum magnifica
tion for any given minimum resolvable spot. Thus, if the spot size is one
millimeter a peak improvement of 59% occurs at magnification of two. Be
tween 1.5 and 3 millimeters, the improvement is at least 43%. In general,
the optimum magnification for any screen unsharpness Uf and focal
spotwidth θ is given by
(Up)
e
Screen brightness varies directly with X-ray intensity. The minimum
perceptible brightness difference due to a change in tire thickness is

/ Phosphor Surface (Film)

UG * Geometric Unsharpness
F= Source Size
t - Object Distance to Recording Surface
D0= Source -Object Distance

FIGURE 8.2.99 X-ray geometry


MEASUREMENT OF TIRE PROPERTIES 645
1 .6 B 38 = — u · AJC where Δx is a change in absorber thickness, u is linear
absorption coefficient for a given material and B is screen brightness in
millilamberts.
Figures 8.2. 100a and b are good examples of a composite X-ray of mis
placed glass belt in a bias-belted tire. The tire had run through one service
life and was prepared for retreading when the buffing rasp removed some
of the belt in the left center of the image. The tire was nevertheless re-
treaded, and failed during a standard wheel test. Note the sharpness not
only of the belt plies but the casing plies at the top of the picture. With
skill, the X-ray technician can identify each of the plies within the tire car
cass. Other characteristic singularities may be found in Figures 8.2.101-
8.2.103.

Ultrasonic
The most recent development in nondestructive tire inspection is the use
of ultrasound. Tire inspection using ultrasound has been developed using
both reflected ultrasonic energy from components within the tire and
transmission of energy through the tire [100], [101]. Each technique has
value for different reasons and these differences will be described, as well
as factors affecting the use of ultrasound in tire inspection.
If it is postulated that separations are a serious drawback in a tire, then
transmission ultrasound is reasonable in spite of the fact that there is not
nearly as much information output as in reflection ultrasound [102].
An example of this is its use in screening used casings for retreading. If

FIGURE 8.2. lOOa. Belt edge location by X-ray


646 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.100b. Marks on lire indicate bell edges.

a tire has used up one service life, it is assumed to have sufficient structural
integrity to be retreaded providing the belts and plys are intact and the
casing has no major cuts, punctures etc., all of which tend to cause inter-
laminar separations. Transmission ultrasound can be used effectively to
sort such tires since separations are found easily.
The reason for the relatively good signal given off by laminar separa
tions has been explained [103]. The sonic energy arriving at a laminar sep
aration is partly reflected and partly transmitted. The energy balance is

where
ET = transmitted energy incident on the separation
EK = energy reflected from the separation
E, = energy transmitted
> through the separation.
The overall transmission of 1 MHz sound has been calculated for thin
films of several materials between tread and carcass rubber. Figure 8.2.104
shows the results of such a calculation for air, for water and for teflon.
It may be seen that if the intermediate material is air, the separation will
be essentially opaque acoustically if the rubber surfaces are separated by
only ten m i II ion t hs of an inch!
The curves for teflon and for water show that if the impedance of the
intermediate material is at all comparable to that of the rubber, the trans
mission is very nearly 100% so long as the intermediate layer is thin com-
MEASUREMENT OF TIRE PROPERTIES 647

FIGURE 8.2.101 Typical non-uniformity of cords


MECHANICS OF PNEUMATIC TIRES

FIGURE 8.2.102 Mismatched bell splices


MEASUREMENT OF TIRE PROPERTIES 649

FIGURE 8.2.103 X-ray of lire bell

Water (r'= 0.733)

Teflon (r'= 1.617)-


'=1.617)
I

Air(r'. 0.00022)

KT 10" 10' to" 10- icr' 10'


INTERMEDIATE LAYER THICKNESS, Mnches

FIGURE 8.2.104 Transmission of I MHz ultrasound through a thin intermediate layer be


tween tread and carcass rubber stock.
650 MECHANICS OF PNEUMATIC TIRES

0
^
u)

FIGURE 8.2.105 Ultrasonic analysis record of buffed lire carcass.

pared to the wavelength. For thicker intermediate layers, maxima and


minima of transmission occur as the layer thickness passes through multi
plies of (λ2/4), the maximum values of reflection remaining small if the
disparity in impedance is not large.
An interesting if unorthodox application of the transmission ultrasound
principle has been in the use of a transducer designed to operate at the
very low frequencies of 25 KHz in air. In frequencies of this range, sound
is transmitted through the cord rubber matrix, and if problems of air rub
ber interface geometry are addressed usable signals may be obtained [104].
Currently devices employing this technique are in the prototype stage
from a number of manufacturers. A typical unit will be a multichannel de
vice displaying its output as a function of rotation. Figure 8.2.105 is one
such trace showing singularities in a buffed casing.
Reflection ultrasound is now beginning to find application in non
destructive inspection of tires in two widely differing implementations.
The first is in use of a single transducer to assess the fatigue life of the ply
structure in truck tires. A study conducted for the U. S. Army [105] has led
to a conclusion that ply structure undergoes progressive degradation
throughout the life of the tire. It is hypothesized that this change can be
monitored by the amplitude of ultrasonic reflections from the ply struc
ture.
At the U. S. Department of Transportation a system is in use which uti
lizes the full capabilities of reflection ultrasound as applied to tire in
spection, as described by Ryan [106]. Figure 8.2. 106a and b show the unit.
The system works on the principle of a three-arm carousel which trans
ports a tire mounted at the end of each arm to one of three stations: load/
MEASUREMENT OF TIRE PROPERTIES 651
652 MECHANICS OF PNEUMATIC TIRES

unload, debubble and inspection. At the load/unload station, a mechani


cal assembly consisting of a retractable pneumatic arm, electromagnets
and an inflation nozzle mounts and inflates the tire. Since the inspection
technique requires liquid coupling, a water bath is provided. At the de-
bubbling station water jets force bubbles off the tire surface to insure suf
ficient "wetness." The tire is then scanned at the inspection station by a
multiple array of sequentially pulsed transducers while it is rotated
through one revolution. The ultrasonic scanning system utilizes an array
of 25 transducers mounted for full tire coverage. The system displays all
25 scan segments simultaneously on a TV video display. This is then re
produced by using a hard copy unit.
Figure 8.2.107 is a photograph of a typical display from a tire. Essen
tially the display is oriented as if the tire were cut across at the serial num
ber, laid out flat and the beads greatly separated by stretching. Each trans
ducer channel is represented by a vertical band which contains the
information for one 360° rotation of the tire. The left-most band is data
from the transducer nearest the whitewall bead. Center channels are in the
tread area and the right hand channel is nearest the blackball bead. The
white lines within each channel represent sound reflected from the lamina
within the tire. Ply splices show up as a system of bumps from one bead to
the other in a diagonal line for bias tires and straight across for radials.
Tire runout is seen as a deviation from straightness of the vertical white
stripes in the case of the tire shown. There is a separation caused by a
piece of paper inadvertently built into the tire. This is shown and an ex
panded photo of the defect is included.
Nondestructive inspection has found application in detecting builders
defects and process malfunctions as a part of a sampling program. Gener
ally this activity has followed the availability of equipment in other fields.
For example, x-ray is routinely used in examination of welded structures.
But generally the technology has been slow to penetrate the field of non-
metalics because of radical variations in bulk properties of the media. It is
anticipated that as the technology becomes available and as increasing de
mands are made on the excellence and uniformity of tires, nondestructive
inspection will play a larger part in tire engineering.
MEASUREMENT OF TIRE PROPERTIES 653

FIGURE 8.2.107 Typical reflection ultrasound display from a tire with built-in defects.
654 MECHANICS OF PNEUMATIC TIRES

References
This bibliography refers frequently to V.R.L.D. reports made by the Vehicle Research
Laboratory at Delft. These reports are in the Dutch language and can be obtained in micro
film or blueprint on payment.
[1] Milliken, W. F., and Whitcomb, D. W., Close, W., and Muzzey, C. L., Symposium on
research in automobile stability and control and tire performance, IME, Auto. Div.,
p. 107; SAE Trans., 344 (1956).
[2] Cough, V. I:,, and Roberts, G. B , Dunlop cornering force machine, Trans. I.R.I.
33(5), (1957).
[3] Kollmann, K., New testing machine for the study of tires, Revue Gén. du Caoutchouc
(10) (1959).
[4] Krenipel. G., Experim. Beitrag zu Untersuchungen an Kraftfahrzeugreifen, Diss.
Karlsruhe (1965).
[5] Henker, E., Dyn. Kennlinien von P. K. W.-Reifen, Wissenschaftl.-Techn. Veröffent-
lichungen Automobielen H3 (1968).
[6] Bruinsma, F., Design and measurements of an airbearing for a running belt with an
electro-hydraulic vibrator system, V.R.L.D. Report No. 472a,b,c (Sept. 1968).
[7] Bird, K. D. and Martin, J. F., The Calspan tire research facility: Design, development
and initial test results, S.A.E. Paper 730582 (May 1973).
[8] AHena, P. H. van., Design of an air bearing, V.R.L.D. Report No. 763 (June 1975).
[9] Spaink, G. N., Friction and drainage measurements on abrasion paper of different
qualities, V.R.L.D. Report No. 391 (April 1967) (Translated by Cornell Aero
nautical Laboratory, Inc., Buffalo).
[10] Van Eldik Thieme, H. C. A., Experimental and theoretical research on mass spring
systems, FISITA 1960 (Elsevier Publishing Company, Amsterdam) p. 386.
[11] Gough, V. E., and Whitehall, S. G., Universal tire test machine, FISITA 1962 (Inst.
Mech. [Ing., London).
[12] Nordeen, D. L., and Cortese, A. D., Force and moment characteristics of rolling tires,
SAE Paper No. 713A (June 1963); SAE Trans., 325 (1964).
[13] Nothstine, J. R , and Beauvais, F. N., Laboratory determination of tire forces, SAE
Paper No. 7138(1963).
[14] Van Eldik Thieme, H. C. A., The measurement of tire characteristics, De Ingenieur,
No. 28 and 30(1964).
[15] cimn, J. L., and Marlowe, R. L., Road contact forces of truck tires, SAE Paper No.
670793 (1967).
[16] Heesewijk, A. P. C. van, and Groeneweg, H. H., Description of six component tire
tester, V.R.L.D. Report No. P055 (Aug. 1967).
[17] Savkoor, A. R., Dynamic behaviour of an orthogonal tyre force measuring unit,
V.R.L.D. Report No. P029 (July 1963).
[18] Edema, L., Measuring platform, V.R.L.D. Report No. P017 (Nov. 1961); P017a (Jan.
1962).
[19] Van der Zee, P., Calibration of platform, V.R.L.D. Report No. 323 (Sept. 1964).
[20] Van Donkelaar, H., Dynamic response of platform, V.R.L.D. Report No. 358 (Sept.
1965).
[21] Yspeert, A. J., Inverse filter for measuring platform, V.R.L.D. Report No. 377 and
377A.
[22] V oermans, J. J., Results of pulsator excitations of measuring platform, V.R.L.D. Re
port No. 422 (Feb. 1968); Design of new measuring platform for truck tires,
V.R.L.D. Report No. 150 (March 1968).
[23] Bakker, C. C., Normal force and longitudinal force response to vertical axle motions,
V.R.L.D. Report No. 486 (Nov. 1968).
[24] Buis, P., Pneumatic tires, V.R.L.D. Report No. P084 (1967).
[25] Koolhof, F. J. W., Polman, J., and Olland, R. C., Static and dynamic testing of pneu
matic tires. V.R.L.D. Reports No. A032, P004-P005 (1958/59).
[26] Cooper, D. H., Radial stiffness of the pneumatic tire, Trans. I.R.I. 40, 58 (1965).
[27] Weber, G., Theorie des Reifens mil ihrer Auswirkung auf die Praxis bei hohen Beans-
pruchungen, A.T.Z. 56(12), 325 (1956).
[28] Stulen, J. H. B., Comparison of road and drum measurements, V.R.L.D. Report No.
326 (Nov. 1964).
[29] Molen, E. L. van der., Measurements of free radius, loaded radius and effective radius,
V.R.L.D. Report No. P172 (July 1974).
[30] Chiesa, A., and Tangorra, G , The dynamic stiffness of tyres, Revue Gén. du
Caoutchouc 36(10), 1321 (1959).
[3 la] Rasmussen, R. E., and Cortese, A. D., Dynamic spring rate performance of rolling
tires, SAE Paper No. 680408 (1968).
MEASUREMENT OF TIRE PROPERTIES 655

[31b] Potts, G. R. and Csora, T. T., "Tire Vibration Studies: The State of Art", Tire Science
and Technology, TSTCA, Vol 3, No. 3, Aug 1975, pp. 196-210.
[31c] Sekula, P. J., Hall, G. L., Potts, G. R., and Conant, F. S., "Dynamic Indoor Tire Test
ing and Fourier Transform Analysis", Tire Science and Technology, TSTCA, Vol.
4, No. 2, May 1976, pp. 66-85.
[32] Krempel, G., Experiraenteller Beitrag zu Untersuchungen an Fahrzeugreifen, Diss.
(1965); A.T.Z., 1 and 262 (1967).
[33] Davisson, J. A., Design and application of commercial type tyres, SAE Paper No. SP-
344 (1969).
[34] Single, C. E., Wide base tires—a new concept for light trucks, SAE Paper No. 680083
(Jan. 1968).
[35] Timan, D. A., and Rooney, J. H. M., Static load deflection relationships of several
tires, V.R.L.D. Report No. P108 (Jan. 1969).
[36] Essers, E., and Kotitschke, J., Ueber die dynamischen Radlasten von Lastkraftwagen,
Techn. Hochschule Aachen (1957).
[37] Vegter, T., Measurements of dynamic wheel loads, V.R.L.D. Reports 214 (1959) and
U020(1960).
[38] Milhlfeld, A., Ein hochfrequenztechnisches Verfahrens für Reifen und Schwingungs-
messung, Diss. T. H. Braunschweig (1949); A.T.Z., 147 (1953).
[39] Bombard, F. J. von, Verfahren zur Messung der dynamischen Radlast beim Kraftwa-
gen, Diss. T. H. Milchen (1956); Deutsche Kraftfahrtforschung 131 (1959).
[40] Burgman, E. F. M., Dynamic tire deflection measurements, V.R.L.D. Report No. 390
(1967).
[41] Knight, S. J., and Green, A. J., Deflection of a moving tire on firm to soft surfaces,
SAE Trans. 5, 116(1962).
[42] Gengenbach, W , and Weber, R., Neues Verfahren zur gleichzeitigen Bestimmung der
Einfederung und der Verformung eines Reifens in Umfangsrichtung wahrend des
Betriebes, Automobil Industrie 3, 93 (1969).
[43] Svenson, O . Untersuchung dynamische Krattc zwischen Rad und Fahrbahn, deutsche
Kraftfahrtforschung 130 (1959).
[44] Senger, G., Ueber dynamische Radlasten beim Ueberrollen kurzwelliger Unebenhei-
ten durch schwere luftreifen, Deutsche Kraftfahrtforschung 187 (1967).
[45] l.ippmann. S. A., Nanny, J. D., Analysis of the enveloping forces of passenger tires,
SAE Paper No. 670174 (Jan. 1967).
[46] Missel, H., Research on tire behaviour on a flat road, Diss. Techn. Univ. Delft (1932).
[47] Floor, W. K. G., The effective rolling radius of pneumatic tyred wheels, Nat. Aeron.
Res. Inst. Report 428 (1954).
[48] Vickers, H. H., and Robison, S. B., Measurement of tread motions and application to
tire performance, Proc. Int. Rubber Conf., 1959.
[48a] Schuring, D. J., Bird, K. D., and Martin, J. F., "Power Requirements of Tires and Fuel
Economy", Tire Science and Technology, TSTCA, Vol. 2, No. 4, Nov. 1974, pp.
261-285.
[49a] Evans, R. D., Factors affecting the power consumption of pneumatic tyres, Proc. Sec
ond Rubber Tech. Conf, London, 1948, p. 438.
[49b] Schuring, D. J., "Energy Loss of Pneumatic Tires under freely Rolling, Braking and
Driving Conditions", Tire Science and Technology, TSTCA, Vol. 4, No. 1, Feb.
1976, pp. 3-15.
[49c] Glemming, D. A. and Bowers, P. A., 'Tire Testing for Rolling Resistance and Fuel
Economy", Tire Science and Technology, TSTCA, Vol. 2, No. 4, Nov. 1974, pp
286-311.
[50] Fogg, A., Measurement of aerodynamic drag and rolling resistance of vehicles,
FISITA Congress, 1964 (Soc. Auto. Eng. of Japan, Tokyo).
[51] Roberts, G. B., Power wastage in tires, Proc. Int. Rubber Conf., Nov. 1959.
[52] Frolov, L. B., and Khromov, M. K., Delerminalion of the rolling resistance of tyres by
an electric torque meter, Sov. Rubber Tech. (9) (1965).
[53a] Williams, T., Power consumption of tyres. Transport and Road Research Laboratory.
TRRL Supplementary Report 192 UC. (1975).
[53b] Clark, S. K., Dodge, R. N., Ganter, R. J., and Luchini, J. R., "Rolling Resistance of
Pneumatic Tires", Univ. of Mich., Report DOT-TSC-74-2, July 1974.
[54] Stiehler, R. D., and Steel. M. N., Power loss and operaling temperature of tires, Proc.
Proc. Int. Rubber Conf., Washington, Nov. 1959, p. 73.
[55] Collins, J. M., The relevance of elastic and loss moduli of lyre components to tyre en
ergy losses. Paper to Div. Rubber Chemistry, Am. Chem. Soc. (1964).
[56a] Kainradl, P., Kaufmann, G., and Schmidt, F., Zusammenhang der Erwarmung von
L.K.W. Reifen mil den visco-elastischen Eigenschaften der verwendeten Gum-
miqualitaten, Kautschuk und Gummi-Kunststoffe (I), 27-36 (1966).
656 MECHANICS OF PNEUMATIC TIRES
[56b] Walter, J. D. and Cunant. F. S . "Energy Losses in Tires", Tire Science and Tech
nology, TSTCA, Vol. 2, No. 4, Nov 1574, pp. 235-260.
[57] Yurkovski, B., Effect of design parameters on the rolling resistance of tyres, Sov. Rub
ber Tech. (1 1), 32-54 (1966).
[58] Khromov, M. K . and Bruev, E. V., Rolling losses of radial ply tires in a wide speed
range, Sov. Rubber Tech. (1%S).
[59] Anonymous, Tyres, a review of current constructions, developments, performance.
Automobile Engineer, 274-288 (July 1969).
(60) Paish, M G . Effects of tractive effort on rolling resistance and slip of pneumatic tyres.
Motor Industry Research Association M.I.R.A. Report 1965/16.
[61] Grosch, K. A , The effect of tyre surface temperature on the wear rating of tread com
pounds, J.I.R.I. 1(1) (Jan. 1967).
[62] Spelman, R. II., Determination of passenger tire performance levels—high speed, SAE
Paper No. 690508 (Chicago, Illinois, May 1969).
[63] Richey, G. G., Hobbs, R. H , and Stiehler, R. D., Temperature studies of the air in a
truck tire, Rubber Age 79, 273-276 (1956).
[64] Homing, V. J., The temperature measurement at the Government Tire Test Fleet,
Rubber Age 74, 395-3% (1953).
[65] Coddington, D. M., Marsh, W. D., and Hodges, H. C, New approach to tire durability
testing, Rubber Chem. Tech. 38(4), 741-756 (1965).
[66] Ludwig, G., Rhodes, D., and Simson, B., Thermistor takes temperature of running
tire, Nat Bur. Stand. (U.S.), Tech. News Bull. 52(6), 119-120 (1968).
[67] Parker, R. C., and Marshall, P. R., The measurement of the temperature of sliding sur
faces, Proc. IME 148, 209 (1948).
[68] Minkes, S. . Surface temperature measurements by an infrared radiation method with
indiunj-antimonide cells, V.R.L.D. Report No. 538a,b,c (1969).
[69] Hoogen, E. v.d., Possibilities to trace hot spots in the tread of tires. V.R.L.D. Report
P012 (1960).
[70] Buis, P., Braking measurements on pneumatic tires, V.R.L.D. Report No. 313 (1963).
[71] Holmes, K. E., and Stone, R. D., Tyre forces as function of cornering and braking slip
on wet road surfaces, R.R.L. Report LR254 (Road Research Laboratory, 1969).
[72] Davisson, J. A., Basic test methods for evaluating tire traction, SAE Paper No. 680136
(Jan. 1968).
[73] Meades, J. K., The effect of tyre construction on braking force coefficient, R.R.L. Re
port LR224 (Road Research Laboratory, 1969).
[74] Horz, E., Die Einfluss von Brerask raf treglern auf die Brems- und Fiihrungskraft eines
gummibereiften Fahrzeugrads, Deutsche Kraftfahrtforschung 195 (1968).
[75] Bajer, J. J., Proposal for a procedure for evaluating wet skid resistance of a road-tire-
vehicle, SAE Paper No. 690526 (May 1969).
[76] Kelley, J. D., Factors affecting passenger car tire traction on the wet road, SAE Paper
No. 680138 (Jan. 1968).
[77] Radt, H. S., The mechanism of tire thump and roughness., SAE Paper No. 332C.
VJ78] Van den Berg, J. and Tromp, E. B., Tire runout measurements (in Dutch). Vehicle Re
search Laboratory P 178, Delft, The Netherlands, 1974.
V [79] Walker, J. C. and Reeves, N. H., Uniformity of tires at vehicle operating speeds. Tire
Science and Technology TSTCA, vol. 2, no. 3, Aug. 1974.
Hutch, C., Gormish, K. J., Corcoran, D. A., High speed uniformity machines and na
ture of tire force variations. SAE 730691 (1972). XlfK)^
Hamburg, J. and Horsch, J., Reduction of tire nonuniformities by machining tech
niques. SAE 710089 (Jan. 1971). Ai \-> \
[82] Lindenmuth, B. E., Tire conicity and ply steer effects on vehicle performance SAE
740074 (Febr. 1974).
Topping, R. W., Tire induced steering pull. SAE 750406 (1975).
Pottinger, M. G. Ply steer in radial carcass tires. SAE 760731 (Oct. 1976).
Klamp, W. K.. and Meingast, J., Higher orders of tire force variations and their signifi
cance. SAE 720463 (1972).
[86] Marshall, K. D. and St. John, N. W., Roughness in steel-belted radial tires—measure
ment and analysis. SAE 750456 (Feb. 1975).
[87] Nedley, A. L., Effects of wheel nonuniformities on the tire-wheel assembly and the ve
hicle. SAE 680005 (Jan. 1968).
[88] Neill, A. H. Jr. and Kondo, A., Correcting vehicle shake. Tire Science and Technology
TSTCA, vol. 2, no. 3, Aug. 1974.
[89] Hofelt, C., Uniformity control of cured tires. SAE 690076 (1969).
i [90] Nordeen, D. L. and Rasmussen, R. E., Factors influencing the measurements of tire
uniformity. SAE 650734 (1965).
. SAE J 332 a. Recommended Practice Testing machines for measuring the uni
formity of passenger car tires. SAE Handbook 1976.
MEASUREMENT OF TIRE PROPERTIES 657
[92] , Messung der Gleichförmigkeit von Luftreifen (Uniformity measurements of
tires). Wirtschaftsverband der deutschen Kauischukindustrie Leitlime 109 (1971).
[93] Caulfield, R. J. and Higgins, R. J., On-car tire grinder for improved ride smoothness.
SAE 720465 (May 1972).
[94] , SAE J 1060. Subjective rating scale for evaluation of noise and ride comfort
characteristics related to motor vehicle tires. SAE Handbook 1976.
[95] Vogel, P. E., Proceedings of the Second Symposium on Nondestructive Testing of
Tires, 1-3 October 1974, DOD/NTIAC, U. S. Army Materials and Mechanics Re
search Center, Watertown, Mass. 02172.
[96] Holographic Techniques for Nondestructive Testing of Tires, April, 1972, NHTSA 72-
4, NT 1SP82 14258-$4.50.
[97] Potts, G. R., "Application of Holography to the Visualization of Tire Vibrations",
1972 SESA Spring Meeting, Cleveland, Ohio, May 23-26.
[98] Vogel, P. E. J., U. S. Army Materials and Mechanic Research Center, Watertown,
MA. "State of the Art of Nondestructive Testing of Tires", Oct. 1973 AMMRC PTR
73.9.
[99] Radiography in Modern Industry (Eastman Kodak Co., 1969).
[100] Halsey, G. H., "The Nondestructive Testing of Passenger Tires", Scientific Testing
Laboratory, Indiana, PA., Paper presented to the 1967 National Convention Society
for Nondestructive Testing, Cleveland, Ohio, Oct. 8, 1967.
[101] Morris, W. E., et al, "Ultrasonic Method of Tire Inspection, Review of Scientific In
struments," Research Laboratories Goodyear Tire & Rubber Co., Vol. 23, No. 12,
pp. 729-734, December 1952.
[102] Gregory, R. K., "Nondestructive Inspection Techniques for Aircraft Tires," April
1968, Southwest Research Institute ASD-TR-68-11.
[103] Ryan, R. P., "Feasibility of High Resolution Pulse Echo Techniques for Automobile
Tire Inspection," June 1973, NTIS PB 231201.
[104] Bessler, H. H., Bobo, S. N., Lourenco, M. J., Wade, W. R., "Nondestructive Testing
System for Retreads," Nov. 1975, Final Report. NTIS No. PB247-083.
[105] Kraska, I. R., "Ultrasonic Inspection for Tire Retreadability, November 1974. Report
on Contract # DAAE07-73-C-10107, U. S. Army Task Automotive Command,
Warren, Michigan.
[106] Ryan, R. P., "A Semi-Automated Pulse-Echo Ultrasonic System for Inspecting Tires,"
March 1977, DOT HS802-104.
658 MECHANICS OF PNEUMATIC TIRES

8.3 Cornering and Camber Experiments


H. C. A. van Eldik Thieme
8.3.1 Introduction
When an externally applied side force F acts at the centerline of a non-
rolling wheel, the lateral factional forces exerted by the road surface on
the tire will cause a lateral tire deformation as shown in figure 8.3.1. The
top part T of the tire, not in contact with the road, is seen to be almost un-
deformed. Usually we are not so much interested in the deformation of a
standing tire as we are in the phenomena of the rolling tire. Therefore con
sider a side force acting on a rolling wheel. The undeformed top part T of
the tire rolls gradually into the deformed condition by coming into contact
with the road surface [1].1
A top view of the lateral deformation is given in figure 8.3.16, assuming
that these lateral tire deformations have attained steady state. At the in
stant of contact with the road, we observe that the peripheral carcass line
already has a lateral deflection A′A. Passing through the contact zone, the
lateral carcass deflection increases until point E is reached, after which
side slip of the line reduces the lateral deflection. Due to some extra lateral
tread rubber deformation EB. the tread elements first coming in contact
with the road adhere to the road surface over the distance AB. The equa
torial line running along the middle of the tread surface shows the lateral
deformation of the tire tread, as represented by line ABD, where all defor
mations are exaggerated for clarity (fig. 8.3. 1c).
The direction of travel of the wheel is indicated by line AB, making an
angle with the plane of the wheel. The angle is called the slip angle. The
resultant lateral tire force f-\ acts a distance t behind the geometric center
C of the contact area, causing the tire to generate a self-aligning torque
Mz. The distance t is called pneumatic trail. The diagram shown in figure
8.3. Id is a schematic representation of figure 8.3.1b. A slip angle α is
formed between the direction of travel of the center of tire contact C and
the x-axis situated in the wheel plane.
Side forces may be due to centrifugal action when cornering, road un-
evenness, small movements of the wheels for directional control, wind
forces, and the lateral component of the weight on a cambered road sur
face. For illustration, some typical cases of automobile response to side
forces will be discussed below.
Nonsteered Vehicles
Consider the simplified case of a vehicle having constant forward speed,
with a driver holding the steering wheel in a fixed straight ahead position.
The vehicle is allowed to change its direction of motion under the action
of lateral forces. This condition has been represented in figure 8.3.2.
The externally applied lateral force must be in equilibrium with the lat
eral forces Fy acting on the tires by the road. These forces result in slip an-
1 Figures in brackets indicate the literature references at the end of thii section.
MEASUREMENT OF TIRE PROPERTIES 659

(II

Equatorial (In*

Whatl plant

Sid. Kip

M.
rFu

FIGURE 8.3.1. An externally applied side force causes a lateral tire deformation.
The slip angle a of the direction of motion with the plane of the wheel is shown together with the resultant lateral tire
force Fv and the self-aligning torque .W

FIGURE 8.3.2. The effect of an externally applied lateralforce on thefront and rear wheel slip
angles.
MECHANICS OF PNEUMATIC TIRES

Forct
Pith of
vthiclt

Resultant
tirt forct
FIGURE 8.3.3. Path of the vehicle at equal front and rear slip angles (neutral steer).

gles «, and a, of the front and rear wheels. The relation between lateral
force and slip angle is governed by the tire characteristic. We first consider
the particular case in which, at equal front and rear slip angle (αf = αr), the
resultant lateral tire force acts in the center of gravity C,;. When the ex
ternal lateral force is also applied at the CG (for instance on a cambered
road), the vehicle drifts sidewards along a linear path, with an angle be
tween its longitudinal vehicle axis and the direction of motion. When this
translatory motion (without yaw velocity) takes place as shown in figure
8.3.3, the automobile is said to have a neutral steer character. A concept
often employed is the neutral steer-point, or the neutral steer-line in the
X-Z plane, upon which an externally applied lateral force will not produce
a yaw velocity. In the particular case of neutral steer considered above, the
neutral steer-point coincides with the center of gravity.
In figure 8.3.4 the response in case of understeer and oversteer has also
been shown. In these cases the neutral steer-point is located behind and in
front of the center of gravity, respectively. For an external lateral force
acting at the center of gravity we have with an understeer vehicle αf > αr
and with an oversteer vehicle αf < αr.
As discussed before, the externally applied lateral force may also be due
to side winds. The resultant side wind force Fw is supposed to act at the so-
called "center of pressure" Cn and this point may be located in front of
the center of gravity CG. The location of C,, depends on the aerodynamic
styling of the vehicle, on the forward vehicle speed, and on the magnitude
and direction of the wind velocity.

nderstter

Neutral steer

Oversteer

FIGURE 8.3.4. Illustration of understeer, neutral steer and oversteer.


MEASUREMENT OF TIRE PROPERTIES 66 1

It is clear that depending upon vehicle inertia, tire characteristics and


magnitude and point of application of the external lateral forces, the slip
angles of the front and rear wheels may differ considerably.
A simplified explanation may be given using an idealized model, where
the center of gravity of the vehicle is assumed to lie in the road surface. It
is furthermore assumed that at front and rear, the left and right wheels
have equal slip angles, which means that the wheels may be imagined as
being compressed together along the axle center lines, and consequently
that the automobile is replaced by a two-wheeled model.
Consider a side force applied at the center of gravity when the vehicle is
moving forward along a straight path. Assume the front wheel to exhibit
larger slip angles than the rear wheels, as shown in figure 8.3.5. This tends
to turn the vehicle away from the applied lateral force F. The rotation
about the instantaneous center gives rise to a centrifugal force Fc which
opposes the applied side force F.
If the side force F produces slip angles αf < αr, then the resulting cen
trifugal force Fc tends to help the side force F. This condition has been
shown in figure 8.3.6.
Consider again a nonsteering vehicle moving forward at a constant
speed of 100 km/hr. in a straight path, but now suddenly subjected to a
side wind with a lateral velocity Vw = 15 m/sec. resulting in a side force
Fw acting at the center of pressure CP (fig. 8.3.7a). It may be of interest to
show some experimental results obtained [2].
It is observed that after a sudden wind force application the transient
motion may develop in a complicated manner. Figure 8.3.7 illustrates such
a behavior, where first (fig. 8.3.7fe) both tire forces oppose the wind force
Fw and afterwards the tire forces appear to change in sign. The variation
of the yaw velocity observed is also shown (fig. 8.3.7d).
Steered Vehicles
Consider a vehicle running through a curve with a transverse slope of
the road surface. The forces acting on the simplified vehicle with rigid sus
pension system are shown in figure 8.3.8.
The equation of forces in the direction of the slope reads:
W V2
Fri + F# = — ' -jT cos Y* ~

Path of
vehicle

af>or

FIGURE 8.3.5. Path of the vehicle when a, > a.


662 MECHANICS OF PNEUMATIC TIRES

Path of
vehicle

FIGURE 8.3.6. Path of the vehicle when af«tr.

When the component of the inertial force Fc is in equilibrium with the


component of the weight (equilibrium speed), there are no lateral tire
forces Fy required in the road surface and consequently the slip angles of
the wheels are zero.
A similar condition arises when going through a flat curve with a ve
hicle speed approaching zero (fig. 8.3.9). In this simplified case of a vehicle
having an ideal Ackermann steer geometry, the wheel base angle ft is
equal to the reference steer angle δref. For R » l, where l is the wheelbase,
the geometric radius of turn Rgeom = l/ft may be taken equal to R. For K—→
0 we consequently have:

The idealized two-wheeled model also shown in figure 8.3.9 can be used
effectively to interpret experimental tire data for application to a vehicle
traveling at a speed V. The steer angle δ required to keep a vehicle on a
constant radius path in a so-called steady state turn is a function of the
front and rear slip angles.
Assuming that all angle:, and traction forces are small, the equation for
the required steer angle follows from figure 8.3.10

P= = 5,^-0,+ a,

5ref - (a, - a,)


Consequently when:
a, = a, then R = /?ieom (neutral steer),
(a, - a,) > 0 then R > Rteom (understeer),
(a, - a,) < 0 then R < Rvma (oversteer).
This simplified analysis of vehicle behavior illustrates the definition of un
dersteer and oversteer in the linear representation.
To obtain a better understanding of the behavior of vehicles at low and
high lateral accelerations, including the influence of suspension system, it
is necessary to consider a number of other important quantities such as
MEASUREMENT OF TIRE PROPERTIES 663

8
B
wind
fore*
B ii

©
V . 100km /h const
Vw:15 m Ittc .

®
0,2 0.4 0,6 B,l 1,0

FIGURE 8.3.7. Path of the vehicle with the resultant side windforce Fw acting in the center of
pressure Cr.

moments of inertia, lateral and longitudinal load transfer, steering system


elasticity, roll steer and roll camber, self-aligning torques at the front and
rear wheels, etc.
It is obvious that the study of directional vehicle control is very com
plicated, and beyond the scope of this book. The reader is referred to the
literature for a more detailed treatment of the subject of vehicles having
664 MECHANICS OF PNEUMATIC TIRES

\
FIGURE 8.3.8. Forces acting on a vehicle running through a banked curve.

steered front wheels [3-6]. Readers are also referred to the reference list of
section 9.5 [29-43].
It is hoped that the simplified introduction given above has shown the
importance of obtaining experimental tire data, such as slip angle, corner
ing force and self-aligning torque relations, to be treated in the following
subsections.

8.3.2. Cornering Experiments


Photography of Contact Area Deformations
The difference in deformation between a vertically loaded free rolling
tire (α = 0) and a cornering tire can be illustrated by photographing the
contact area of tires through a glass plate with grids during slow motion
[5] (fig. 8.3.65).

georri

geom

FIGURE 8.3.9. The reference steer angle &,# and the geometric radius of turn Rf
MEASUREMENT OF TIRE PROPERTIES 669

FIGURE 8.3.10. Idealized two wheeled model.

Measurements of the resulting deformations can be carried out in the


same way as previously discussed in section 8.2.5 for the case of a braking
tire.
First a photograph is taken of the free rolling tire (α = 0). Then a sec
ond photograph can be taken of a standing laterally undeformed tire, set
at a predetermined angle to the direction of motion of the movable plat
form, representing static condition.
Afterwards, using the same negative, a third photograph is made on top
of the second one, but now with the tire rolled in the laterally deformed
steady state cornering condition (fig. 8.3.11). The difficulty with this
method is the experience necessary to synchronize the photographs, but it
is also very time consuming to analyze the photographic data to define the
form of the equatorial line.
A glass plate device to study tire deformations is also used at several
proving grounds by driving a vehicle straight onto the glass plate with the
front wheels toed-in to a predetermined steer angle. Due to steering sys
tem flexibility it is difficult to measure the exact slip angles, and it seems
desirable to use a slip angle measuring device as described in section 8.3.3,
figures 8.3.43-44.
The photographic technique is either to attach the camera onto the ve
hicle and photograph through the glass plate, looking at the contact patch
in a mirror, or to photograph from beneath as tires pass over the glass sur
face [7].

Movable Platform Measurements with the Gough Apparatus


A. Lateral deformations
The deformation due to the lateral flexibility of a tire rolling at a slip
angle may also be obtained with a flat surface modified version of the
original Gough apparatus [6] mounted in the flat platform machine [5]
(fig. 8.3.65). The platform travels at a constant speed of about 0.05 m/sec.
666 MECHANICS OF PNEUMATIC TIRES

^Stf^-

front

Direction movement
of platform

Front

FIGURE 8.3.1 1. Cornering experiment showing deformation of equatorial line.

When the platform has travelled such a distance that the lateral tire de
formations are assumed to have reached steady state condition, the tire is
allowed to roll over the Cough apparatus [10]. The original version of this
apparatus, mounted behind the moving glass plate, had a steel upper sur
face. In order to avoid a difference in coefficients of friction on steel and
glass, both surfaces are now covered with a sheet of transparent perspex.
As discussed [5], the Cough apparatus has three measuring bars.
The bar measuring the lateral force distribution is equipped with a steel
fork which can move in a narrow slot, situated on the lateral centerline of
this transverse bar of the Cough apparatus. Lateral sliding of the tread
element penetrated by the steel fork, with respect to the platform (fig.
8.3.12), is recorded using a potentiometer coupled to the steel fork [10].
We speak of sliding because a relative motion of a tread element with re
spect to the road surface is observed.
MEASUREMENT OF TIRE PROPERTIES 667

-»»- Movement of platform.


FIGURE 8.3. 1 2. Movable platform measurements showing lateral tire deformation and sliding
of a tread element with respect to the platform.

Because the test tire is mounted on a structure with a turntable having a


transverse slide, the mechanism allows for a lateral movement of the
wheel center plane with respect to the platform. The initial lateral defor
mation A′A of the tire where contact begins (fig. 8.3.1b) can now be taken
into account by a suitable transverse movement CM of the slide, so that
the steel fork penetrates the tire tread surface exactly at point A of the

fork movement

Equatorial lint

Wheel center plant

FIGURE 8.3.13. Geometry of lateral slip and deformation of the equatorial line within the
contact zone.
668 MECHANICS OF PNEUMATIC TIRES

equatorial line as marked on the tread (fig. 8.3.13). This is done by obser
vation in the longitudinal direction of the platform at the position of the
steel fork with respect to the equatorial line.
At low slip angles we observe that a tread element coming into contact
with the slowly moving platform first adheres almost fully to the surface,
so that the assumption of no lateral sliding is reasonable ( Vsx = Vsy = 0)
and the original undeformed equator line travels parallel to the direction
of motion. The steel fork does not move laterally.
As the tread element moves further in the contact zone, it produces a
deflection v with respect to the wheel plane which, depending on the lat
eral carcass stiffness Cc and the rubber tread element stiffness [27], in
creases linearly with increasing distance, causing an increasing lateral
force until the local limiting factional force is reached. The tread element
then begins to slide towards the wheel center plane. This lateral sliding of
the tread element with respect to the platform is measured with the mov
able steel fork.
We are interested in the displacements u and ν of a tread element with
respect to its position in the undeformed situation in the X and Y direction
respectively.
The initial lateral displacement ν of the tread element A, which first
comes into contact with the platform (fig. 8.3.13) can be found to within a
small error, using the assumption that A′A" = 0. Hence, as represented in
figure 8.3.13,
ν = CM cos α - (AM - CM sin α)tan α.

The vertical force measuring bar gives a reference signal on a record in


dicating the moment that contact begins with the platform. This is point A
on the equator line, M being the center of the turntable supporting the test
tire, and is indicated by a microswitch signal received from the moving
platform, so that the distance AM is known.
The distance CM is read from a scale attached to the support of the lat
eral slide mechanism, and the rotation angle a of the turntable is also read
from a scale.
By transformation of the fork measurements, as indicated, the lateral
deformation v can now be calculated.

B. Longitudinal deformation
As already discussed in section 8.2 (see [5] of sec. 8.2 and fig. 8.3.63), the
measurement of the longitudinal deformations with a camwheel mounted
in the longitudinal force bar is a very time consuming procedure because
several runs are necessary to obtain a reasonable accuracy. Due to the lat
eral sliding, other neighboring tread elements will come into contact with
the circumference of the camwheel. However, the track of one individual
tread element of the equatorial line over the surface can be constructed
approximately because the lateral movement of one tread element, as
measured with the fork, is known.
By repeating the longitudinal deformation measurements several times,
by giving the slide a lateral movement of 5 mm. or less after each run, the
real longitudinal slip line of one tread element can be reconstructed [9]. As
can be seen in figure 8.3.14, several lines are drawn parallel to the direc
tion of travel of the platform. The different points of intersection of these
MEASUREMENT OF TIRE PROPERTIES 669

lines with the deformed equator line give an indication of the position
along the contact length of the original tread element. The real longitudi
nal slip line, is now reconstructed with parts of the individual longitudinal
slip lines belonging to the points of intersection with the lateral slip line as
shown in figure 8.3.15.
C. Combination of lateral and longitudinal deformation
The lateral and longitudinal movement of a tread element of the equa
torial line can now be calculated and figure 8.3.16 shows the paths of a
tread element, parallel and perpendicular to the direction of motion, for
different slip angles α [24]. From these results it is seen that without brak
ing or driving forces longitudinal sliding occurs, and that the lateral slid
ing at the end of the contact zone is almost perpendicular to the direction
of motion.
At increasing slip angles the deformation of the contact zone will in
crease, resulting in increasing lateral forces. It is seen that at larger slip an
gles the adhesive zone decreases, resulting in an increasing sliding zone at
the rear of the contact area (fig. 8.3.17).
D. Sliding velocity
As can be seen from the lateral deformation curve of the equatorial line,
a tread element gradually rolls into its deformed situation. We observe at
the leading edge A where contact begins a lateral deflection ν with respect
to the undeformed situation (fig. 8.3.18).
Assuming adhesion over the distance AB, the drifting tire shows an
equator line which is straight and parallel to the vector V, hence - ∂ν/∂x
= α = constant. Because no lateral sliding occurs in this region, the lateral
sliding velocity V = 0. Consequently, the value of the deformation veloc
ity Vα of the tread element will be equal to the lateral component V sin α
of the forward velocity V, so that we can write:
V,y=- Vsina+ Kd»=0.

Direction literal
fork movement
-Direction of
platform motion

FIGURE 8.3.14. Longitudinal deformation measurements in the contact zone.


670 MECHANICS OF PNEUMATIC TIRES

Contact length

. Lateral
sliding

Longitu
> dinal
sliding
signals

Longitudinal sliding
Approximated

FIGURE 8.3. 15. Reconstruction of the longitudinal slip line ofa tread element moving through
the contact zone.

Travelling further in the contact zone over the distance BH, we observe
a decrease of ∂ν/∂x until point H, where we obtain the value ∂ν/∂x = 0
and Vsy = — y sin α.
Due to the change in slope beyond H of the lateral deformation curve of
the equatorial line, the lateral deformation velocity V,, now changes in sign
resulting in further increase of the sliding velocity.
Having reached the rear end of the contact zone at point /), the tread
element of the equatorial line gradually returns to the wheel center plane
as indicated in the figure [24].
E. Lateral force distribution
In discussing the lateral deformation of the tire, we observed for moder
ate slip angles a that a tread element on entering the contact zone at first
adheres to the platform. As the tread element moves further in the contact
MEASUREMENT OF TIRE PROPERTIES 671
SlUing
p»r»U«l (o
direction
oftr*vtl t Tlrt S. 10.13
mm
Load 300 kff
Infl prt»«ur«1,5lcgf/cml

0 ~ S 10 15 20
(mm)
Slidinjj. dirtctlon of tr.vil

FIGURE 8.3.16. Trend element paths as a function of slip angle.

zone, it produces a deflection which increases linearly with increasing dis


tance, causing an increasing lateral force. Dependent on the vertical force
distribution, which gradually drops to zero at the rear of the contact zone,
the local limiting friction is reached, and the tread element begins to slide
back towards the wheel center plane, thus reducing the lateral force, as
shown in figure 8.3.18. Note that the peak value of lateral force distribu
tion is not reached until after sliding has started.
When the tire travels over the lateral force measuring bar of the Gough
apparatus, we obtain the lateral force function. The resultant lateral force
/•', may be obtained by integrating the recorded force function, and can be
compared with the total lateral force Fy, as measured with the six-com
ponent tire tester, also mounted in the movable platform machine.
For moderate slip angles a, as shown in figure 8.3.19, the resultant lat
eral force /•', acts at a distance t behind the contact center C. The distance t
is known as the pneumatic trail.
Neglecting the longitudinal force distribution in the contact area, we
find, in this simplified case of purely lateral slip, a value for the moment
about a vertical axis Mz = — Fν · t.
In reality the longitudinal forces in the contact area influence the mo
ment about the contact center C (fig. 9.5.40b), resulting in a total corner
ing moment Mz known as the self-aligning torque. The force Fy is usually
called the cornering force. A more complete set of curves of the lateral or
cornering force distribution over the contact area, obtained with the
Gough apparatus, are given for various slip angles in Chapter 5 (fig. 5.58).

Direction of traval Sliding zone

FIGURE 8.3.17. The sliding zone at the rear of the contact area.
672 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.3.18. Distribution offerees, lateral sliding distance and lateral velocity over the
contact length.

Lattral Fore*
Distribution

FIGURE 8.3.19. The resultant lateral force Ff acts a distance t behind the contact center C
(pneumatic trail).
MEASUREMENT OF TIRE PROPERTIES 673

The result of the cornering force-slip angle relation is shown in Chapter 5


(fig. 5.59), also indicating the pneumatic trail.
Bars for measuring the lateral force distribution have also been built
into a drum of 3 m. in diameter, and the results obtained are similar to
those just discussed [10].
F. Tread wear resistance
A nondestructive estimate of resistance to tread wear [11] of various tire
constructions is very important to the tire manufacturer, because tread
wear tests on the road require a large number of tires and are usually quite
time consuming.
Although laboratory studies of movement of tread block elements on a
glass plate add to the knowledge of wear, they fail to solve problems of
varying rates of average wear.
The original Gough apparatus [8] made it possible to give an estimate of
work done by the lateral factional forces during passage through the con
tact patch when cornering.
The force as measured with the lateral force measuring bar is recorded
against the lateral movement of the central rib of the tread. The resulting
diagram gives, by measuring the area under the curve, the work done by
the lateral force when cornering. The tests are made at a range of slip an
gles. From results obtained, it is found that the work done plotted against
the cornering force coefficient FJFN are straight lines for different loads
when shown on log-log paper.
Tread wear measurements obtained from road tests, made with vehicles
driven continuously around a constant radius at constant speed, showed
similar straight lines when tread wear was plotted against the cornering
coefficient on log-log paper. Neglecting longitudinal slipping, it follows
that work done by the lateral forces; is strongly related to tread wear. The
fact that the slope of the lines on the log (work done) versus log (cornering
coefficient) is influenced by the tire construction (fig. 8.3.20), means that
some decision must be made as to what cornering coefficient should be
chosen for the basis of comparison of work done. The choice was a corner
ing coefficient of 0.3, which corresponds to a vehicle executing a lateral ac
celeration of 0.3 g. [11].
The average value of 0.3 g. has also been obtained from statistical accel
eration studies with 9 different makes of cars and 6 drivers, where the lat
eral component of the gravitational force [12] due to the roll angle has also
been taken into account.
The figure indicates that the work done, and consequently the wear rate
of a conventional bias tire is higher than a radial belted tire under similar
conditions.
Road and Indoor Test Measurements
A. General observations
Although thousands of road and drum tests have been made it is felt to
be outside the scope of this chapter to discuss the influence of parameters
such as type of cord, cord angles, number of plies, rim width, aspect ratio,
stiffening of sidewalls and reinforcement of beads, bias, bias belted, radial
belted, or combinations of radial bias tires, etc. Due to its frictional char
acter, the cornering force curve is also dependent on the texture of the
road surface, the rubber compound and curing, the temperature and the
speed of sliding in the contact patch. It will be obvious that the data men-
674 MECHANICS OF PNEUMATIC TIRES

Work done
log scale

I
08

02

(09 tcilt
0,3 as
Cornering coefficient » _JL

FIGURE 8.3.20. Estimate of work done by the lateralfrictionalforces during passage through
the contact zone when cornering.

tioned above are not always available, and therefore parameters which
can be influenced by the vehicle engineer will be discussed, such as slip
angle, normal load, inflation pressure and speed.
f As indicated, tire cornering force and self-aligning torque data depend
on so many factors that a general discussion is very difficult.
However, some attention will be given in this section to the influence of
tire construction, tread pattern and tread resilience For further informa
tion the reader is referred to chapter 6 and its literature references.
Further it should be noted that in order to obtain reproducible results, a
standard test procedure which conditions the tire is required.
The data presented here for /•'. and Mz are obtained under steady state
rolling conditions. When a tire is made to roll at a slip angle α, it gradually
rolls into its deformed shape as in figure 8.3.1, thereby developing an in
creasing lateral force with increasing distance travelled. After a distance of
about 4-5 times the contact length the lateral deformations nearly attain
their steady state values, and the lateral force /•', will remain constant (fig.
8.3.2 1a).
Recently, increasing attention has been given to modeling vehicle han
dling in emergency maneuvers, stimulating interests in tire dynamic prop
erties.
Figure 8.3.21b illustrates a typical lateral force amplitude ratio as a
function of slip angle path frequency as obtained by the Calspan's Tire
MEASUREMENT OF TIRE PROPERTIES 675

Lattral FM. 300 kgf


Forct X Pi.1,5kgf/cm2
kgf a. *•
200 radial belttd

.X convention] t bias

100 ^*^

7 1.0 Distance travelled (m)

FIGURE 8.3.21a. A tire made to roll at a constant slip angle a develops an increasing lateral
force with increasing distance travelled and attains the steady stale value.

Research Facility [40], compared with other measurements and calcu


lations.
The dynamic response of a tire to inputs of vertical load, slip angle and
camber angle is always delayed. Especially for accident avoidance maneu
vers, tire characteristics obtained under steady state conditions will not be
sufficient to predict the vehicle motions. Schuring [40] gives results of an
extensive test program to obtain the dynamic response characteristics of
car tires.
The sideforce response to a rapid slip angle change is:

where Fy· . „ = the steady state side force


where tr = relaxation time, time to attain 63% of Fy

0.5 1.0 10
<u, , rod /ft

FIGURE 8.3.216. Lateral force amplitude ratio ,, as a function of slip angle path
frequency u
676 MECHANICS OF PNEUMATIC TIRES

Generally a quarter revolution of a tire is needed to attain 63% of the Fy.^


At one revolution the static side force value is almost achieved.
B. Effect of slip angle α
a. Dry roads
The influence of the slip angle a on the lateral force Fy for constant nor
mal load FN and constant inflation pressure /?, is shown in figure 8.3.22.
For small values of the slip angle of approximately 1-2 degrees, the lat
eral force function F,.(«) is usually considered to be linear.
The slope of the curve indicates the so-called cornering force stiffness
CFα = (∂Fy/∂α)α=0. This value is very much dependent on the tire construc
tion due to different values of the lateral carcass stiffness c,. and tread rub
ber stiffness cp (fig. 9.5.33). Because it may be observed that at the larger
slip angles the lateral force f\ increases, but the adhesive zone decreases,
resulting in an increased sliding to the rear of the contact zone (fig. 8.3.17),
the shape of the Fy(α) relation may be understandable.
The laterally stiffer tire reduces tread motion in the contact patch, and
thus gives less wear. The tire designer can influence the shape of these
curves considerably by increasing or decreasing the cornering force stiff
ness. It should be observed that the recorded Fy(α) curves, as shown in Fig.
8.3.22 very often do not pass through the origin. It is a common practice to
shift all curves horizontally so as to pass through the origin, in order to
compare results, thereby neglecting the well known existence of cornering
forces at zero slip angle.
The same shifting applies for the self-aligning torque curves shown in
the following figures.
As shown in figure 8.3.19, the resultant lateral force Fy acts at a distance
t behind the contact center, partly explaining the value of the cornering
moment Mz. At increasing slip angle α the lateral force /', increases, result
ing in an increasing moment Mz. The slope of the Mz(α) curve, as illus
trated in figure 8.3.23, indicates the cornering moment stiffness CMa «
Due to the decrease of the pneumatic trail t with increasing slip angle,
the moment Mz reaches a maximum, and then drops to lower values.

Fz = 3.3 kN
Pi =175 kN/m1
size 165-13
5.90-13

a - steel belted radial


b-glasstibro bias-belted
c - textile belted radial
• d - cross ply

^-slipangle (deg)

FIGURE 8.3.22 Lateral force vs. slip angle for 165-13 tire.
MEASUREMENT OF TIRE PROPERTIES 677

Road dry Road dry


asphaltic concrete
KC asphaltic concrc
Tirt 5.60-13
V = 10 km/h
' •
2tO * F. > 300 kgt

-b Tin 5.60 13
200 V . 1.0 km/h
S F, m 300 kgt
110
. 1.7 kgt /cm
"'
120
P
M
(0
/'
10 12 U
-Slip U a

Tires a ami 1' are of different construction

FIGURE 8.3.23. Effect of tire construction on lateralforce and self aligning torque versus slip
angle.

The moment Mz = ∫A(px · y~ P,' x)dA is Influenced by the longitudinal


and lateral force distributions, which are asymmetrical in shape over the
contact area A.
We observe that the rolling resistance force is now acting out of the
wheel plane due to the lateral tire deformation. The local forces developed
are also dependent on the longitudinal and lateral carcass and tread stiff
nesses. Dependent on the vertical force distribution in the contact area,
the local limiting frictional force may be reached and sliding occurs when
^Jp\ + p* = up,. But as observed in section 8.3.5 (fig. 8.3.52), the coefficient
of friction /i, depending very much on the speed of sliding, reaches a maxi
mum at a certain sliding velocity and then drops to lower values. In the
rear zone of the contact area the lateral sliding velocity Vsy increases (fig.
8.3.18), which may result in a decrease of the coefficient of friction.
Combining all these effects, the shape of the Mz(α) curve may now be
explained, and also the fact that a positive self-aligning torque M z can be
obtained at high slip angles α.
To indicate what happens when a dry road becomes wet, the F.,(α) and
Mz(α) relations are shown for both conditions for another tire (fig. 8.3 24).
b. Wet roads
The cornering characteristics can change remarkably on wet roads hav
ing different road textures as illustrated in fig. 8.3.25a. As shown in the fig
ure some aspects have opposite effects. For instance a worn tire attains a
Jvigher Cf. value but a lower /*„ value [41].
"The difference in behavior on dry and wet roads may be best illustrated
under extreme conditions, such as on a wet smooth polished asphalt test
track. This road, as measured with the drainage meter (sec. 8.3.5, fig.
8.3.55) showed zero drainage, and a low skid resistance value of about 25,
as measured with the Portable Skid Resistance Meter (sec. 8.3.5, fig.
8.3.57), indicating the dangerous situation of no micro- and no macro-
roughness. The wetted road temperature was about 18°C. The test results
were obtained with the towed trailer (see ref. [1] of sec. 8.2, also fig. 8.2.8)
by increasing the slip angle a very slowly, under constant vertical load FN
and at a constant speed of 40 km/hr.
678 MECHANICS OF PNEUMATIC TIRES

_j
C*«fficieM Tl
* ^*^ —•— .. f
f
V^
*fy ™»d "j
w* * —7**
~A2
fir ^^***'
• 1 "m ™" Tlrt tOO.U
^ Speed 30 km/h
/

7 10* W 20*
F • 400 kg!

Slipinglt

FIGURE 8.3.24. Effect of road condition on the cornering force coefficient and self aligning
torque-slip angle relationships.

It is observed on wet roads that the lateral force Fy increases with in


creasing angle α (fig. 8.3.25), but drops at larger slip angles. Returning to
lower slip angles, it is seen that the lateral force /', cannot recover its ini
tially higher values as shown by the closed loops at the indicated speed of
40 km/hr. At a slightly lower speed (30 km/hr.), this loop has not been ob-

Tl m 330 kqt
Pi . 1.8 bar
tire 165-13
til
Fz ^-^

I
£-1 t tl^57max= (J?^ _ ^]7^
f smooTh surface rough surface
^fFoT
0 A. .. .j. '»«" ^ptwu low speeo
—•» slipongle oc worn tire full tread depth
low Cpm high CF<M

75 kgf/deg
cross ply tire steel belted radial worn
full tread depth

FIGURE 8.3.2Sa. Characteristic values of lateral force coefficients for different road surfaces.
MEASUREMENT OF TIRE PROPERTIES 679

Fire 5.60.13
V .40km h
F..300 kgf

FIGURE 8.3.256. The effect oftire construction on the corneringforce-slip angle relationships.

served in testing tire "a" (fig. 8.3.26). The variation in lateral force for tire
"b" is less than for tire "a," indicating a slight preference for tire "b" un
der these conditions.
An explanation of this phenomenon is that possibly a coefficient of fric
tion, ju, arises which depends upon the orientation of the tread element rel
ative to the road surface. The orientation is related to the lateral deforma
tion of the tire and consequently to the cornering force Fy . The function of
wet friction coefficient versus sliding speed may abruptly change to an al-

Tire = 5.60-13 (a)


F,,. 270 kgf
Pi »1.7kgf/cm2

FIGURE 8.3.26. The effect of speed on the cornering force-slip angle relationships.
680 MECHANICS OF PNEUMATIC TIRES

ternative representation when a critical value of the cornering force Fy is


passed. The self-aligning torque values Mz for tires "a" and "ft" are also
different as shown in figure 8.3.27ft. The footprint area of both tires has at
equal loading conditions the same value, but the value of the contact
length of tire "a" is about 1 5 percent higher. This may explain the higher
maximum value of the self-aligning torque of tire "a" compared with the
lower value of tire "ft."
The tread pattern, tread resilience, and tread hardness, obviously play a
significant role. Difference in skidding performance may be largely attrib
uted to these differences in tread pattern design and resilience. The center
rib of tire "a" has, contrary to tire "ft," no sipes or kerfs.
Both tires have four drainage grooves, but the width of the drainage
grooves of tire "a" are smaller than those of tire "ft," and the outer ribs of
tire "ft" have more sipes. Tire "ft" has a lower tread rubber resilience than
tire "a" and the tread rubber hardness of tire "ft," with a value of 61, is
higher than tire "a," the latter having a hardness of 56.
To judge which tire is better from this test alone may be very difficult
because the normal load (fig. 8.3.28a, ft) also plays an important role when
cornering, due to weight transfer.
The phenomenon that the lateral force coefficient cannot always be re
covered after reversal of the slip angle may be of great importance to ve
hicle handling performance. It indicates that under certain conditions of
speed and friction, vehicle break away may not be readily recovered by
decreasing the steering wheel angle.

C. Effect of normal load


The influence of the normal load W = FN on the cornering force Fy is
shown in figure 8.3.29. In discussing the vertical load-deflection curves
(sec. 8.2.4, fig. 8.2.49), we observed an increasing contact length with in-

Rbad wet
—•
polished asphalt
Tire 5.60.13
V =40km/h

FIGURE 8.3.27. The effect of tire construction on the self aligning torque-slip angle
relationships.
MEASUREMENT OF TIRE PROPERTIES 681

smooth, polished asphalt


Tire = 5.60- 13 (a)
V =tdkm/h
P -1.7kgf/cm2

Road wet ^
shed asphalt
Tire 5.60.13(b)
V = 40km/h
, .1.7 kgf/cm2

FIGURE 8.3.28. The effect of normal load on the cornering force-slip angle relationships
shown for lire a and tire b.

creasing load at constant inflation pressure, resulting in almost no increase


of the vertical contact pressure/^ (ch. 5, fig. 5.32). The cornering force of
ten shows at each slip angle a certain maximum value. The observed drop
of the lateral force f1, at increasing load /•'.. and at larger slip angles may
be explained by the influence of the increasing longitudinal forces and the
decreasing lateral stiffness. The longitudinal force distribution shows in
creasing forces with increasing load (fig. 8.2.49) and therefore influences
the lateral force /•', obtainable. Combining this effect and the decrease in
682 MECHANICS OF PNEUMATIC TIRES
Cornering Force
Dry asphalt J
FV ffsoo- Tire 5.60 - 13
kgf Speed 30 km/h
\i p. = 1.5 kgf /cm3

Normal load kgf

FIGURE 8.3.29. Influence of load on cornering force

lateral stiffness with decrease in normal load, the shapes of the lateral
force curves may be explained.
Due to the nonlinear shape of the Fy-FN curve (fig. 8.3.30), load transfer
from one wheel to another has an effect upon the values of the slip angles
required for the generation of the total side force which balances the cen
trifugal force Ff when cornering.
If the Fy-FN curve were linear no loss in lateral force Fy would be ob
served at constant slip angle. The load sensivity is a measure of how much
the tire is able to increase the lateral force produced at one degree slip
angle as the load is increased. It is evaluated between 0.8 and 1 .0 times the
rated load [42]. Because the weight of the car remains the same when cor
nering, load transfer results in increase of the required slip angles a to bal
ance the centrifugal force f',.. A similar effect is observed when driving on
bad, uneven roads with large variations in the normal load, also causing a
loss in cornering force due to the nonlinear shape of the curve (cf, sec.
9.5.2). The load transfer sensitivity is a measure of how much total lateral
force is lost by a pair of tires when one is reduced in load by a certain
amount from a base load value and the other is increased by that same

2 F,

r*i

FIGURE 8.3.30. The loss in lateral force /•', due to load transfer.
MEASUREMENT OF TIRE PROPERTIES 683

amount. This is evaluated at four degrees slip angle and between 0.4 and
1.6 times the tire rated load [42].
The self-aligning torque Mz increases with increasing normal load due
to the increasing contact length, as shown in fig. 8.3.31 [15].
The Gough plot of the Fy-Mz relation for different slip angles at con
stant normal load and inflation pressure is shown in figure 8.3.32.
Truck tire characteristics
Because the literature gives relatively little information about the me
chanical properties of truck tires, it may be of interest to show some results
as obtained on a flat bed tire testing machine at a speed of 2 km/h [43].
Corntrlng Force
Fy Kgf 7M

Normal load kgf

Road dry asphalt


Tir« MO.U
Stlfaligning Torque Speed 30 km/h
-Mi .1.4 kgf /cm2
Kgf m

FIGURE 8.3.3 1 . Cornering force and self aligning torque versus normal load.
684 MECHANICS OF PNEUMATIC TIRES

ROAD-DRY ASPHALT
Tire 8.00-14
Speed 30km/hr
FN>400kgf
Pi • l.4Kgf/cm*

FIGURE 8.3.32a. Gough plot of cornering force versus self aligning torque.

The carpet plot of lateral force versus tire load and slip angle as shown
in figure 8.3.32b illustrates the characteristics of a 12.00-20/G truck tire at
80 psi inflation pressure. The cornering stiffness Cα versus tire load of
three types of truck tires is shown in figure 8.3.32c. Increasing the tire load
causes an increase in contact length and an increase in lateral stiffness. A
tire showing higher cornering stiffness will develop more lateral force than
a lower stiffness tire at the same slip angle and normal load.
D. Effect of inflation pressure pi
Increase of inflation pressure results in an increase of the lateral stiff
ness, as observed from static load-deflection measurements (sec. 8.2.2, fig.
8.2.31) and therefore the lateral force Fy tends to increase at constant nor
mal load and constant slip angle [16].
The self-aligning torque Mz decreases with increasing inflation pressure
/>, because the contact length decreases, resulting in a decrease of the pneu
matic trail t.

16° 9900
Tire
Lood(lb)
_
tr
o
4200

2100

FIGURE 8.3.326. Carpet plot of lateral force vs. tire load and slip angle.
MEASUREMENT OF TIRE PROPERTIES 685

15-22.5/H
(90 psi)

2000 4000 6000 8000 100000


FIGURE 8.3. 32c. Cornering stiffness for three types of truck tires.

Results obtained from measurements with a drum of 4.0 m. in diameter


[17] are illustrated in figure 8.3.33a.
Results obtained for a 10.00-20G truck tire on a flat bed machine at a
very low speed of 2 km/h are shown in figure 8.3.33b for a rated inflation
pressure of 100 psi and at 50 psi. As may be observed a reduction of the
inflation pressure at high load levels will strongly influence the lateral
force. The cornering stiffness exhibits similar pressure sensivity at higher
vertical loads [43].
E. Effect of tread wear, tread pattern
As the tire wears the tread pattern as well as the tread depth are
changed. Particularly for truck tires the tread pattern grooves are nar
rower with decreasing tread depth. Kerfs are often just a few millimeters
deep. Especially on wet roads the traction characteristics can change re
markably. For truck tires the full tread depths are generally from 12 to 14
nun. The stiffness of the rubber blocks and ribs in the tread pattern plays
an important role in the cornering characteristics on dry roads, figure
8.3.33c. The cornering stiffness increases to about 150% compared with a
new tire when the tread depth decreases to zero as shown in Table 8.3.1.
For the worn condition the tread pattern plays a role too. A rib type
tread pattern has a relatively high tread compliance compared with an
open tread pattern as is usual for traction or mud and snow tires. The
open pattern has approximately twice the void area of the closed rib-type
pattern. This can result in remarkable differences in longitudinal stiffness,
cornering stiffness and camber stiffness [43].
In table 8.3.2 are the characteristics parameters shown for three differ
ent tread patterns, two rib-type and one open (M&S) tread pattern. The
values were measured at rated inflation pressure, 85 psi, and rated load
5430 Ib.
F. Effect of velocity
The influence of speed on the lateral force / ', is considered to be rather
unimportant on dry roads. The effect of increased speed may be a slight
increase in lateral force. The added tension component in the carcass due
to the centrifugal force tends to increase the tire stiffness, but the contact
length decreases.
686 MECHANICS OF PNEUMATIC TIRES

Tin 5.20-13
FN=250kgf
V±40km/h

F, 300- -

100

Normal load

FIGURE 8.3.33o The effect ofinflation pressure pt on selfaligning torque and corneringforce.

On wy{ rqads. however, the decrease in lateral force f„ may be consid


erable at higher slip angles 14]. The tread pattern and state of the road
surface are extremely important parameters determining the directional
stability. Although the speed dependence of the cornering force of pat
terned and smooth tires has been shown in chapter 6, figure 6.30, it may be
of interest to show some cornering force coefficient (C.F.C.)—-slip angle
relations in figure 8.3.34 as obtained on different wetted road surfaces with

Detail of Test Surfaces


No. Description Texture
1 9.5 mm quarzite macadam surface rough, harsh
4 9.5 mm Bridport macadam surface rough, polished
5 mastic asphalt smooth, polished
MEASUREMENT OF TIRE PROPERTIES 687

6000 1000-20G
design lood
1005430
psi Ib

FIGURE 8.3.33*. Lateralforce versus slip angle and vertical load on 10.00-20/G tire at rated
pressure (100 osit and at 50 osi.

patterned and smooth bias ply tires, the depth of water being of the order
of 0.5-0.75 mm.
The patterned tire was size 5.25-16 and had a Lupke pendulum resil
ience value of 31 at 20°C, compared with a smooth 5.00-16 tire having a
resilience value of 55. The hardness values were approximately the same,
namely 62 and 63, respectively.
The details of the test surfaces are given in the table below, and the
numbers of the test surfaces are the same as in figure 8.3.34 [14].
The influence of speed and surface is clearly illustrated, and it is shown
that the initial part of the curve is nearly linear and independent of the
surface, except at the higher speeds. The patterned tire gives greater
C.F.C. values than the smooth. In general, tires of low tread resilience give
greater values than those of high resilience tread rubber. Therefore the
smooth tire, having a high resilience tread rubber gives extremely low
C.F.C. values on the smooth polished mastic asphalt surface No. 5.
In the extreme case of aquaplaning conditions, the cornering forces fall
to such a low value that a vehicle may be directed from its straight path by
1QOO-20F
FullWom
85psi HolfWom
New

FIGURE 8.8.33c. Lateralforce versus slip angle and vertical load on a 10. 00-20/F tire in three
states of wear.
688 MECHANICS OF PNEUMATIC TIRES

TABLE 8.3.1 Measured Mechanical Propertiesfor Nylon 10.00-20/F Rib-Type II Tire


Design in Three States of Wear
New Half Worn Fully Worn
C5, Ib/unit slip 42,000 52,000 60,000
C., Ib/deg 523.4 690.5 771.5
Cr Ib/deg 69.0 101.4 147.7
Ky> Ib/in 1,618 1,784 1,866
Kz, Ib/in 4,700 3,939 4,600
C.—longitudinal stiffness
Ca—cornering stiffness
Cy—camber stiffness
Ky—lateral spring rate
1C,—vertical spring rate

the influence of a sidewind or a cambered road. The feel at the steering


wheel should give sufficient early warning, but in certain cases the self-
aligning torque—cornering force relations can have a distorted shape (fig.
8.3.35), suggesting that the steering feel may not always be a reliable in
dicator [18].

8.3.3 Camber and Cornering


Camber
The inclination of the wheel plane to the vertical indicates the camber
angle γ and is called positive when the wheel leans outward at the top and
negative when it leans inward (fig. 8.3.36).
The undeformed top part of the tire rolls gradually into the deformed
condition by coming in contact with the road surface, as illustrated in the
figure. A lateral tire deflection is observed, resulting in a camber force F, ,.
As demonstrated in figure 8.3.37 the lateral force Fyγ developed by a
camber angle γ is much smaller than the lateral force Fn developed by an
equal slip angle α. The results shown are measured on a drum of 4.0 m. in
diameter [17]. Similar relations obtained on a drum of 2.5 m. in diameter
are shown for another tire (fig. 8.3.38).
For this conventional bias tire the camber force developed can be ap
proximately related to the normal load by:
Fr - FN tan y = CFT • y
with a constant camber force stiffness CFγ, because of the linear character
of the lateral force—camber angle relation.

TABLE 8.3.2 Measured mechanical properties of 10.00-20/F nylon tire in three tread
patterns. A-rib-type l:B-rib-type II: C-open tread
(a) Rib-type I (b) Rib-type II (c) Open Tread
c. 46000 42000 28000 Ib/unit slip
c. 508.2 523.4 516.0 Ib/deg
>CT 56.7 69.0 39.9 Ib/deg
Ky 1477 1618 1291 Ib/in
K. 5032 4700 4500 Ib/in
MEASUREMENT OF TIRE PROPERTIES 689

F,.230kgf

FIGURE 8.3.34. The effect of velocity and tread pattern on the cornering force coefficient-slip
angle relationships on various wet roads.
Surface I (rough, harsh), surface 4 (rough, polished), surface 5 (smooth, polished).

SPEED ,O mile/n SPEED SO mile/n

s£ 0-10
z
g
"0 O01 O02 0 001 OO2
SELF ALIGNING TORQUE/LOAD -(t

- - - - - tyre B
A Satisfactory tread pattern.
B Unsatisfactory tread pattern characteristics.

FIGURE 8.3.35 The effect ofspeed on the corneringforce-selfaligning torque relationships on


a wet road.
690 MECHANICS OF PNEUMATIC TIRES

F..-F,

FIGURE 8.3.36 The camber angle y produces a lateral tire deformation resulting in a camber
force Fn.

Radial ply tires show less camber force. The resultant lateral camber
force is said to act a certain distance ahead of the contact center, and this
distance is called the pneumatic lead, resulting in a camber moment Mzγ.
The camber moment is usually small and may be neglected. The ex
planation of the camber moment is given by considering the rolling tire to
consist of two narrow tires, mounted rigidly a distance 2b apart on a
spindle. As can be seen in figure 8.3.36, the rolling radius r1 > r2. Because
the distances travelled are equal, but the rolling radii are different, an anti
symmetric longitudinal slip must occur, producing two equal longitudinal
forces Fxy, which act in opposite directions, resulting in a moment Mz =
2F^-b (See also fig. 8.5.21).
Combination of Camber and Cornering
The combination of lateral forces due to a slip angle a and a camber
angle γ is shown in figure 8.3.39, and according to the sign convention it is
shown that for positive values of α and γ the lateral forces F,,t and Fn act
in the same directions. The influence of the camber angle γ on the total
lateral force l\ decreases with increasing slip angle α, due to sliding in the
contact area. This is best illustrated (fig. 8.3.40) at low normal load, for the
case where F^ and Fn both act in the same direction.
A positive camber angle γ develops a camber force at zero slip angle,
which can be counteracted by giving the wheels toe-in.
Static toe-in of a pair of wheels is the difference in the transverse dis
tance between the wheel planes taken, respectively, at the extreme rear
and front points of the tire treads. When the distance at the rear is greater
the wheels are said to be "toed-in" [3].
Camber angle variations due to change in wheel track of independent
suspension linkages result in lateral sliding or "scrub" over the road sur
face. The scrubbing action can result in unacceptable tire wear.
The camber angle variation with respect to time causes the wheel to be
subjected to a gyroscopic moment Mz = Ix · £2 · γ, where Ix is the polar mo
ment of inertia.
MEASUREMENT OF TIRE PROPERTIES 691

Tir« 5.20-13
kgf
V - 40 km/h
240
p. « 1,6 kgf /cm2

200

160

120

•0

iO

Tir. 5.20-13
V . 40 km/h
PI • 1,6 kgf /cm2
a . o*

300 400 kgf


Normal load F_

kgf
60 Tirt 5.20-13
> 300 kgf
SO P| '• 1,6 kgf /cm2 ^
a > 0' ^
40

^
20

<
0
^
2* 48 6' »-
8* Camber10*angli y

FIGURE 8.3.37. Comparison of cornering force and camberforce values.


692 MECHANICS OF PNEUMATIC TIRES

Camfctr fere*

NorcMl

Tlrt U0.13
Camktr ftret
kgf
70

50

»•

Camktr anflt V

FIGURE 8.3.38. Camber force versus normal load and camber angle.

Actual S/Jp a/irf Camber Angle Measurements


A well known method of obtaining slip angles is by using a vehicle with
two symmetrically toed-in front wheels. However, elastic deformations
due to braking forces can influence the slip angles introduced into the sys
tem. Because the tire forces of a vehicle under steady-state cornering con
ditions are also of interest, another method is required to measure the slip
and camber angle of the test wheel.
Assuming zero camber angle γ, the unknown slip angle a can be found
by tracing the path of two spring loaded ballpoint pens, situated a fixed
distance / apart on a subframe of the axle stub [19]. The pens will trace
two lines separated by a distance d, depending on the slip angle α in
troduced as in figure 8.3.41. It follows that sin α = d/l. If the direction of
travel coincides with the jc-axis of the platform (Ψ = α), the cornering
force Fy is obtained from the road platform tire tester (sec. 8.2.1):
Fy = — FH sin a + Ff cos a
where Fs and F9 indicate the forces on the tire.
A better method is using two electronic flashlamps triggered about 100
times per second with a short duration of the flash to ensure reasonably
clear photographs at different vehicle speeds. The flashlamps are placed
MEASUREMENT OF TIRE PROPERTIES 693

FIGURE 8.3.39. The combination of lateralforces due to camber and cornering.

T.rt .520-13 Tin =5.20-13


V < tOkm/h V =iOkm/h
Y = const Y -conlt

F. = ISOkgl FN > ISOkgl


^ sl.Skgt/cm1

FIGURE 8.3.40. The effect of normal load on the total lateral force Fy due to camber and
cornering.
694 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.3.41. Slip angle measuring apparatus mounted on a wheel passing over the road
platform tire tester.

on a distance / on the axle stub in holders showing a small rectangular slot,


giving dashed lines on a photograph. The photo apparatus is placed in a
structure well above the road (figs. 8.3.42a, b and 8.3.61).
A small inclination of the axle stub arm due to cambering will hardly
influence the accuracy of the slip angle measurements. But obtaining the
slip angles from the dashed lines on the photograph is time consuming
work. Therefore this method has been used to control a special slip and
camber angle measuring device (S.C. meter) as shown in figure 8.3.43. The
idea for such a device originated in 1964 with the Motor Industry Re
search Association (M.I.R.A.) in England.
The principle of the slip angle measurement is that a free trailing axle
(1), gives the required direction of motion of the front wheel of a motorcar
(fig. 8.3.44). Because two small spring-loaded trailing wheels are used, the
tubes (2) of the supporting frame are always perpendicular to the road sur
face. The joint (3) between axle stub and frame allows the measurement of
the camber angle. The joint (3) is situated in the vertical center line of the
tubes (2).
The slip angle measurements, however, require correction charts as a
function of the lateral acceleration [19, 20].
The result is that the S.C. meter can be used on smooth and slippery or
even on bumpy roads as a tool to measure with reasonable accuracy the
required angles under cornering conditions.
Of course, this device can also be used without the road platform (fig.
8.3.43). As an example, this device gives information about car "waddle"
when driving at low speeds on straight flat roads, due to tire nonuni-
formity. The slip angle varies in such a case from 0.1-1.0 degree (sec.
8.2.6).
MEASUREMENT OF TIRE PROPERTIES 695

Photoapparatus

Daih No 1 a 3 4 5
'•of, 01785 0.1475 01183 009118 H0837
'« <*« 0.1538 onifl Ofl96>i QD816 00805
«l , 10' 7' 8* 2V 6%5' 5*25' k'v'
*1 8' 4 5' 8*1.0' 5-30' k'kO' 1-36'

FIGURE 8.3.42. Photographic slip angle measuring technique using two electronicflash lamps
placed on the axle stub.

8.3.4 The Influence of Braking and Traction on Cornering


Dry Surface Measurements
The combination of a longitudinal force Fx and lateral force /•', , gives a
resultant horizontal force K as shown in figure 8.3.45 for the case of a
braking tire.
696 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.4.43. Slip and camber angle measuring device.

The difference in behavior of a bias ply and radial ply tire is illustrated.
As can be expected, the lateral and longitudinal carcass and tread rubber
stiffnesses will greatly influence the results obtained.
The radial ply tire produces a more or less symmetrical shaped curve for
braking and traction. The bias ply tire produces a more pronounced slope
in the curve, as shown in dashed lines.
The bias ply curve illustrates the effect that braking gives a higher ob
tainable lateral force Fy than traction, for a given value of the longitudinal
force /\ , at constant vertical load and constant inflation pressure.

FIGURE 8.3.44. Principle of the slip and camber measurements.


MEASUREMENT OF TIRE PROPERTIES

radial ply
bias ply

FIGURE 8.3.45. The influence of braking and traction forces on the lateral force for a radial
ply and bias ply tire.

The fact that larger traction forces often require larger slip angles a to
obtain the same lateral force f, is illustrated in figure 8.3.46, thus influenc
ing the over- or under-steer character of a vehicle.
Measurements taken on a dry steel drum of 4 m. in diameter are shown
in figure 8.3.47a, b for a number of slip angles [17].
The effect of braking and traction forces on cornering force /•', and self-
aligning torque Mz is illustrated in figure 8.3.48 for two different loads at a
constant speed of 40 km/hr. and constant inflation pressure of 1.4 kgf/cm2
[17].
Wet Surface Measurements
Numerous tests have been carried out since 1960 with the Delft tire test
trailer (fig. 8.2.8), but results have not been published in the literature. The

wheel plane

FlOURE 8.3.46. The effect of traction forces on the lateral force.


698 MECHANICS OF PNEUMATIC TIRES

Tin US-IS ( biai ply)


V-40 km/h
FB-250 kgf
Pi -U kgf/cm»

kgf 200 160 120 SO <0 0 40 80 120 160 200 2*0

FIGURE 8.3.47o. Friction circle results for a bias ply tire.

results obtained on different wet road surfaces with various tires, however,
are similar to those reported by the British Road Research Laboratory
[14], and because these results are readily available, they will be discussed.
The force measurements are made on a fifth wheel which can be set at a
slip angle in the range of 0-20°, and whose angular velocity can be held at
any desired value, independent of the vehicle speed, by means of a hy
draulic transmission driven from the normal vehicle drive [21].
In the following the results obtained with only one tire will be discussed.
For further information the reader is referred to the literature, where de
tailed information is given on the test procedure and the evaluation of re
sults, with further details of 10 test tires and 5 test surfaces [14].
The effect of the combined action of a longitudinal force Fx and a lat
eral force Fy on a wet surface is illustrated in figure 8.3.49 for a patterned

V.40 km/h
F. , 250 kgl
P| > 1,4 kgf/cm*
Tin* 165 -15 (radial ply)

kgf 200 160 120 SO 40 0 40 SO 120 160 200 240 kgf

^BRAKING TRACTION*

FIGURE 8.3.476. Friction circle results for a radial tire.


MEASUREMENT OF TIRE PROPERTIES 699

Drumttst (D-i m«ttr)


Tirt (.SO- 1]
V- 40 km/ti
FH nom ci 300 kg?
F..200 inn 400 kg I

FIGURE 8.3.48. effect of longitudinal forces on the cornering force-self aligning torque
relationships given for two normal loads.

radial ply tire, with steel reinforcing belt. The test surface was a 9.5 mm.
Bridport macadam, having a rough, polished texture. The depth of water
was in the order of 0.5-0.75 mm.
A small braking force, generating little slip, does not affect the corner
ing force stiffness very much, but reduces the maximum cornering force
coefficient (C.F.C.) and the slip angle at which the peak is observed. It is
shown that at increasing braking force coefficient B.F.C., the C.F.C. curve
breaks away and reaches a lower maximum than before. At B.F.C. values

FIGURE 8.3.49. The effect of the combined action of a longitudinal force and a lateral force
on a wet surface is shown for constant brakingforce coefficients (BFC) and constant braking
percent slip KH.
700 MECHANICS OF PNEUMATIC TIRES

greater than the locked wheel value, the curves become closed loops, as
shown for a value of 0.35 B.F.C., illustrating the effect that a large braking
force cannot be satisfied at slip angles greater than a certain amount [14].
The 0.3 B.F.C. curves do not appear as loops, because they extend outside
the slip angle range covered by the figure.
The dashed lines, indicating constant braking slip K,,, are quite different
from those of constant B.F.C. If a definite braking slip is imposed, the
C.F.C. slip angle curve has a smaller initial slope, a lower peak, and at
tains the peak at larger slip angles.
The effect of percent braking slip κB on the cornering force coefficient
C.F.C. at a given slip angle is shown in figure 8.3.50 for three different ver
tical loads. The well known fact that the cornering force falls off very rap
idly with braking slip is clearly demonstrated on this type of road surface.
The effect of brake slip on the braking force coefficient B.F.C., is also
shown in the figure for slip angles of 0, 4 and 10 degrees. These effects are
important when considering anti-locking devices.
The effect of traction, as observed with a special test vehicle, has also
been reported [14], but discussion of results obtained will be omitted.
However, because cornering traction methods, using conventional vehi
cles, may be of interest from the standpoint of vehicle safety and control, a
short description will be given.

Road wet
_-•^—•
Rough. polii hid macadam
Tirt 1(5.15 (Radial ply)
Spud (I km/h

s&
03 v\\ F^.IMkjf 03
° '
y 02 \\\
0.1 vcs
\S
20 (0 20 (0

Braking t\jp p*r cent K-

02

0 10 JO M U 50 M 70 K 90 100
^_ Braking slip pir ctnt K,

FIGURE 8.3.50. The effect of braking percent slip KB on the corneringforce coefficient CFC at
a given slip angle for three vertical loads.
The braking force coefficient (BFC) versus braking percent slip Kg u also shown.
MEASUREMENT OF TIRE PROPERTIES 701

Conventional vehicle tests


Because under maximum braking or driving conditions, the available
lateral force component is almost zero, there will be no force available for
the vehicle control. In cornering traction tests utilizing standard vehicles,
all tires on the vehicle are subjected simultaneously to their individual
complex conditions of load, inclination, torque, slip angle, speed, etc. [22].
The nontethered cornering traction test is based on the combined effects
of all tires on the vehicle and their ability to generate sufficient forces to
overcome the centrifugal effects on the vehicle.
For this test a car is driven in a circle at increasing speeds until break
away occurs. The driver follows a procedure of increasing the speed in
0.5-1 km/hr. increments on alternate laps until a clearly distinguishable
. level of the vehicle control is lost or rear breakaway is detected.
The maximum speed attained is usually the basis for tire comparison.
The momentary road surface condition, water depth, wind, temperatures
of road and tire, etc. play an important role in the maximum attainable
speed.
Out of many procedures one test is to drive the vehicle in third or fourth
gear on the test surface at a predetermined speed while trying to follow a
curved path. Arriving at the curve, the vehicle is accelerated by means of a
full throttle down shift. The driver follows again a procedure of increasing
the speed in 1 or 2 km/hr. increments on successive laps, until at the so-
called trace speed a distinguishable level of vehicle control loss at the end
of the curve is observed. The speed just before this loss of control is called
"justholding." When skidding already occurs at a certain point, we speak
of moderate slip.
Results obtained with a standard vehicle, on a smooth polished wet as
phalt surface, gave for the tires a and b of figures 8.3.25-28 the following
speeds in km/hr.:

Tire a Tire b
Just holding 37.5 40.9
Trace 37.5 42.8
Moderate slip 40.9 45

The large difference in cornering characteristics of tires a and b on the


same test surface have been discussed previously (figs. 8.3.25-28).
The combined effects of accelerating or braking while cornering are also
considered in the so-called slalom test, where the car is driven alternately
around a series of equally spaced markers positioned on the test surface.
Although a variety of test methods have been devised, all methods are
subject to many factors which may adversely affect the repeatability and
accuracy of data. Even when A.S.T.M. E-17 control tires are used for es
tablishing the friction coefficient rating of the test surface, this practice
does not provide the basis for the comparison of test data among different
tests. A program to determine the tractive performance level for all modes
of operation requires a large number of tests, and to reduce variations the
individual tests have to be repeated several times in order to average the
data obtained [22].
702 MECHANICS OF PNEUMATIC TIRES

8.3.5 Difficulties in Measuring Forces and Moments


General Observations
The forces and moments acting between tire and road depend on the
distribution of the local normal pressure and the local tangential stresses
in the contact patch. The tangential stress occurring at a point x, y in the
area A (fig. 8.2.8) is determined by the local normal pressure and the local
friction coefficient between tread and road surfaces. The normal pressure
is dependent upon the construction of the tire and the design of the tread
pattern for a given load and inflation pressure. The coefficient of friction
of a rubber compound against a clean dry road surface varies with the
speed of sliding and the temperature. Under isothermal conditions, the co
efficient of friction tends to a constant value at very low sliding speeds, ap
proaching static conditions. As the speed of sliding increases, the coeffi
cient of friction rises and attains a maximum value at a certain speed (fig.
8.3.5 1a). The friction coefficient at constant speed varies with the test tem
perature, as shown for a butadiene rubber compound on a glass surface
(fig. 8.3.52).
It is generally recognized that in the region of low sliding speeds, the
isothermal friction-speed-curve increases with speed. At a certain speed,
the coefficient μ attains its maximum value. For instance at -25° C a max
imum is reached at about 10-2 m/sec. On further increasing the speed, the
coefficient decreases. The variation of the coefficient with the logarithm of
speed has a bell-shaped form. For speeds exceeding the speed at which the
maximum friction is obtained, we observe a negative slope of the friction-
speed-curve causing stick-slip. Despite the fact that the drive mechanism
of a measuring system imposes a constant speed, the sliding speed of the
surface of the rubber test specimen is not uniform. The tangential force
varies periodically between its maximum and minimum values, as shown
in figure 8.3.51b.
Both the inertia forces and the actual sliding speeds are unknown.
Therefore some authors have taken the average value of the tangential
force to represent the friction force, while others consider the highest value
of the periodic force to be representative. In the equation of motion of the
measuring system, the negative slope introduces a negative damping term.
A simple way to eliminate the stick slip is to introduce a positive damp-

Q01 0.1 1.0 10 100


Sliding speed. V.mph

FIGURE 8.3. 5 \a. The effect of sliding speed on the adhesion coefficient.
MEASUREMENT OF TIRE PROPERTIES 703

r! Sue V*1.2 cm/tec


1 5 sec V'2.10 cm/sec
180 h
120 120 [
60 60 f
0 F r/V
I (tec)

I 5«ec V4.1O
5iec V-0.24cm/iec 180 1—1 /—
120 F *—/
60f * B

Kite) " I dec)


OUASIHARMONIC SAW-TOOTH

FIGURE 8.3.516. Comparison of Force Traces (A) with and (B) Without Stick-Slip. SBR on
Stainless Steel Temp. -45 "C.

FIGURE 8.3.52. The effect of temperature on the coefficient affliction at constant speedfor a
butadiene rubber on glass.
704 MECHANICS OF PNEUMATIC TIRES

ing which is larger than that resulting from the negative slope. The posi
tive damping can be obtained by making use of another friction pair, by
choosing a sufficiently large normal load. The stabilizing friction pair is
mechanically coupled to the test pair, which is housed in a climatic cham
ber. An automated tribometer system for isothermal friction at a speed
range from 10-5 cm/sec to 1 cm/sec for a climate chamber is available in
the Delft Research Laboratory [44].
Other difficulties are that at speeds higher than a few centimeters per
second, the temperature rises above the ambient temperature and it is not
clear which value of the temperature should be considered effective in
governing the friction properties, because the higher temperature occurs in
a layer beneath the surface. Another complication is that friction coeffi
cients are very sensitive to surface conditions, e.g. resulting from con
densation of atmospheric moisture. Only a few points governing friction
laws have been discussed but it may be clear that various factors influence
the factional behaviour of rubber [48, 49].
The Nature of the Track Surface
In order to evaluate tire characteristics obtained from road tests, the
magnitude of the road surface influence should be assessed. When com
paring tire test data it seems advisable to include in the program test re
sults obtained with a special standardized reference tire as a yardstick for
the frictional rating of the road surfaces. If possible the road surface char
acteristics should be obtained separately, in the form of a description of
the macro- and microroughness of the test surface (fig. 8.3.53). Because
the friction of rubber is temperature dependent, the surface temperature
plays an important role and should be given in test reports.
Macro texture
The effective friction on wet roads is controlled by the removal of the
fluid film throughout the contact patch of the tire. The road surface drain
age via macro texture (1-10 mm.) is thought to have an influence upon the
fluid displacement.
A photograph of the surface texture should be made available, in
dicating if required the size of coated chippings or quarzite macadam.
Stereo photography to determine the mean void width has been used,
giving a good picture of the form of the texture [26]. In other cases, profile
traces are obtained with an electromechanical roughness meter.
Stereo photographs have been taken to record the surface profile, from
which the "profile ratio" was evaluated. The profile ratio is defined as the
ratio of the length of the surface profile along a line to the length of the
base line. This method used to assess the profile bears some relation to tex
ture depth and takes into account the shape of the profile. Analysis has in
dicated that the top 40 to 50 thousandths of an inch (1-1.25 mm.) is the
most significant part of the profile in determining the decrease in braking
force coefficient with speed. It has been suggested that the decrease in co
efficient of friction ju,, (κB = 100%) from 50 km/hr. to 125 km/hr. should
not be greater than one quarter, with a minimum value of ft.,, = 0.3, as
measured with the small trailer apparatus of the Road Research Labora
tory [27].
In the sand patch method a known amount <p of fine dry sand or powder
is applied to the surface and is distributed in such a way that the sand just
MEASUREMENT OF TIRE PROPERTIES 703
Surface typ»
(T) Smooth

® Fine textured, rounded


(5) Fine textured, gritty
(7) Coarse textured, rounded
(D Coarse textured, gritty

20 46 60 «0
Sliding spcid.V, mph

FIGURE 8.3.53. The effect of the surface condition on the skid resistance.
(Skid number SN it the ratio of skid resistance to wheel load times 100.)

fills the depressions and hollows in the area A covered. The drainage prop
erty is considered proportional to the ratio <p/A.
The average texture depth (TD) in mm is determined as follows:

TD = 10 -^

<£ = volume of sand in cm3


D = diameter of the area A covered with sand in cm.
This method gives no information concerning the separation and shape
of the texture height and the effectiveness of the surface drainage chan
nels. We are interested in the channelling system and not in the hollows.
We also want to take into account the draping behavior of tread rubber
over the individual asperities reducing the effectiveness of the surface
drainage channels.
With a drainage meter according to a design by Moore [28] the water
removing properties of a surface can be determined. A transparent bot
tomless cylinder is provided on its underside with a rubber ring upon
706 MECHANICS OF PNEUMATIC TIRES

which the meter is placed on the surface to be investigated. The drainage


meter is loaded with the desired number of load rings so that the rubber
ring will drape itself over the asperities much as a tread element does. The
time required for the water level to fall from the top mark to the bottom
mark was originally measured and characterized the drainage potential of
the surface.
Figure 8.3.54a, b shows the Debit drainage meter with an electromag
netic operated valve to fill the outflow meter accurately to the desired
level. In order to measure the drainage on coarse roads it appeared neces
sary to compensate for stones or road unevennesses under the rubber ring
by placing foam rubber between the cylinder and the rubber ring. The real
measuring time was reduced to 10 seconds [29]. In view of the large num
ber of tests, the measurement results are automatically printed with a digi
tal recorder. Other data, such as water temperature and number of load
rings, are fed into a special computer program giving the drainage num
ber, denning the degree of drainage of the macro texture of the road sur
face in question [30].
A new drainage meter measures the drainage as the total water flow
during 10 seconds at different water pressures (0.4-7 kgf/cm.2).
The drainage meter is shown in figure 8.3.54fe. After lowering frame A
from the test trailer of figure 8.2.8 to the road surface, the contact pressure
between rubber ring and road can be varied with the air pressure in cham-

Framc of test trailer

FIGURE 8.3.54a. Road drainage meter.


MEASUREMENT OF TIRE PROPERTIES 707

FIGURE 8.3.546. Road drainage meter.

water IS- 15
08 contact
pressure pressure
kgf/cm1 kgt /cm1
te 10
0.4

as- highway
roadsurface

rough
o 02 04 oe o 02 a* as
{J)v flow dm'/sec y tlow dm'/sec
influence of road surface texture influence of contact pressure

FIGURE 8.3.54c. Influence of road surface texture and contact pressure onflow rate.
708 MECHANICS OF PNEUMATIC TIRES

tin

FIGURE 8.3.54<£ Delft lire tread drainage measuring equipment

her 5 of part B. This part B is then fixed in its position with the aid of a
disk brake (7). The water pressure in chamber 2 is variable between 0.4
and 7 kgf/cm.2.
The device can be used in the laboratory as well as on a test trailer for
measurements on public roads. If required, laboratory measurements can
be carried out on replicas of epoxy resinous material or on samples of the
original road surface.
Results obtained with this drainage meter are shown in figure 8.3.54c
for two different road surfaces. There is a strong indication, that for con
tact pressures in the range of truck tires (6-7 kgf/cm.2) no drainage exists

water
pressure 3
kgt/cm2

I I tread depth 2mm


' HfuU tread depth

i
| 165SR13
load 330 kgf
<J>vH 'n'lj*'ori w kgf/cm1
1 »flow dmj/sec

FIGURE 8.3. 54e. Characteristic values for tread pattern drainage capacity
MEASUREMENT OF TIRE PROPERTIES 709

FIGURE 8.3. 54/. Measurement of drainage capacity of a tread pattern

on the concrete road, whereas for the average passenger car tire pressure
of 1.5 kgf/cm.2 a reasonable drainage is still available.
This influence of the tire contact pressure has also been observed on
towed truck trailer road tests, using the force method and giving the
locked braking force coefficient μls (fig. 8.3.55b).
Another illustration of the influence of the road surface texture and the
influence of the contact pressure between rubber ring and road is given in
figure 8.3.54c.
Because the removal of the fluid film throughout the contact patch of
the tire depends on the drainage capacity of the road as well as the drain
age capacity of the tire tread pattern, a drainage meter has been developed
[45], as shown in figure 8.3.54d. The device consists of a stainless steel box
provided with an orifice (5 X 90 mm) in the top surface. The tire is loaded
on the box, with the orifice in transverse direction. Water is forced from
the orifice through the tread pattern of the test tire. The waterflow and wa
ter-pressure are recorded. As a result, we obtain the characteristics values
('<>•>, </v. and A,,.,, defined as follows (figure 8.3.54e).
Pmax = the water pressure at which the tread in the contact area starts lift
ing from the device.
P0, = defined as 0.9PmK.
<k,9 = waterflow at P0.9.
A09 = characteristic drainage area shown hatched in figure
In figure 8.3.54e are shown two curves (I and II) for a tire with full tread
depth and for a worn tire with an average tread depth of 2 mm. In figure
710 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.3.S5a. The British portable skid resistance tester.

0 20 40 60 BO WO km/h 0 20 40 60 80 WO km/h

FIGURE 8.3.S5b. — Wet skid resistance ranges for car and truck tires.
MEASUREMENT OF TIRE PROPERTIES 71 1

8.3.S4/ a tire can be seen during the test. The meter does not pretend to
simulate exactly what happens to a rolling tire on a wet surface. The sole
purpose is to obtain drainage numbers to compare tread patterns. These
numbers are expected to correlate with wet skid resistance. At the outset, it
was not certain which of the coefficients /*„„, $„,„ or A,,,, correlated best
with the measured values and other tread pattern parameters. Therefore,
all three coefficients were initially taken into account.
These other parameters are the air ratio (AR) and shore length (SL) ra
tio. They are determined from a black ink picture of the footprint of the
loaded tread pattern. The percentage total white area of the total gross
contact area, black plus white, is the air ratio number AR. It usually
amounts to 20-30%. The shore length SL is denned as the total length of
the boundaries of all grooves and sipes, divided by the circumference of
the footprint. The SL value being in the range of 400-1200%.
The results obtained with multifactor test programs emphasizing road
factors and tire factor have been discussed in the literature [41].

Micro texture
Small sharp points in the road surface can penetrate a thin fluid film,
but can also penetrate into the rubber tread surface of the tire. Because the
surfaces have thin films of oxide layer and water, they are far from chem
ically clean and therefore it is not entirely a water film penetration. A mi-
croroughness of 0.05-1.0 mm. on top of the macro texture produces a high
resistance to slip. The adhesive friction process consists of the formation of
adhesive bonds at the real area of contact, at the tips of the hard asperities,
and is caused by the normal load. The elastically stored energy in shear,
due to the tangential force, will try to overcome the surface energy of the
hard solid so as to free adhesive bonds [24]. Only the rubber molecules
forming the real area of contact may be considered near enough to the
field of forces of the hard solid and the deformation is therefore concen
trated in a very thin layer below the surface. Under the action of the tan
gential force the adhesive bonds break and a fresh cycle then begins with
formation of new bonds elsewhere on the surface.
In extreme cases, the stresses at the tips will be large enough to rupture
the rubber, causing abrasion, as has been proved in sliding on dry quar-
tzite [32]. The measurement and classification of the road surface micro-
roughness with a mechanical roughness meter is a difficult subject because
the microroughness is superimposed on the texture.
The small-scale macroscopic roughness is sometimes measured with a
foil-piercing technique. In this technique a piece of aluminum foil placed
on the road surface is given an impact by a rubber tipped rod released
from a predetermined height [33]. The sharper tipped particles pierce the
foil and the number of piercings per square centimeter are counted.
Another more reliable method using replicas has been developed by the
Dutch State Road Building Laboratory. A cast of the surface is made with
synthetic resin of silicon rubber. The cast is sectioned and the section sur
face is projected on photographic paper. With a special optical system the
profile can then be measured and evaluated [25].
Skid testers
The British Portable Skid Resistance Tester is a pendulum device which
measures the friction resistance of a wetted surface to the passage of a rub
712 MECHANICS OF PNEUMATIC TIRES

her slider. Upon release from a horizontal position the pendulum and
pointer swing through an arc, the pendulum returns but the pointer stays
at the farthest point of the arc (fig. 8.3.55a), and the number read from the
scale gives the skid resistance value (S.R.T.). This instrument gives as a
first approximation a reasonable indication of the micro-roughness, but
cannot measure the drainage properties of the road surface. It is a low
speed tester (2.8 m/sec.) and cannot sense the friction level at higher
speeds. As a result, this tester rated the two surfaces number 3 and 5 of fig
ure 8.3.53 as very similar [23], even though the friction levels of these sur
faces were quite different at sliding speeds of 40 mph. The skid number
SN is defined as the ratio of skid resistance to wheel load times 100. A skid
number of 50 implies that the locked tire generates a skid resistance of 50
percent of the wheel load [23].
The shortcoming of all low-speed portable testers, their inability to
sense the drainage properties of a surface and its friction level at higher
speeds, initiated the study and design of several Delft skid testers [34, 35].
Also a mul Itifactor examination of wet skid resistance of car tires has
been conducted. The results can be averaged in a formula for /j. showing
the influence of road texture parameters

In wet countries with mixed traffic, it is vital to safety that there be no


dangerous discrepancy in stopping ability of trucks and busses with re
spect to passenger cars. Since most truck and bus tires operate at high in
flation pressures, when compared with passenger tires, it is interesting to
evaluate truck tire tests versus passenger car tire tests on surfaces having
various macro- and microroughnesses.
With the Delft tire skid trailer [36, 37] an experimental program has
been initiated to study the behavior of truck tires having differences in
tread design, compound and construction on a variety of road surfaces.
Towed trailer road tests results obtained from passenger car tires and
truck tires are indicated in figure 8.3. 55b. The large difference in the
locked braking force coefficient /i,,, for the average passenger car tire and
truck tire is clearly demonstrated [45].
Conditioning of the Tire
Due to surface conditions of both tire and road, and the temperature
dependent character of the friction, a precise description of the test condi
tions is absolutely necessary to compare test results of different tires.
The initial wear of the top layer of the tread rubber of a new tire, and
the warming-up effect during rolling and slipping, produce large varia
tions in the cornering force as shown in figure 8.3.56. These are typical
cornering force-slip angle relations as a function of the number of tests on
a 6.40-13 bias angle tire with a natural tread rubber compound under a
constant normal load of 350 kgf. and a constant inflation pressure of 1.50
kgf/cm2. In test number one, the new cold tire was pressed against a drum
of 250 cm. in diameter, rotating at a constant speed of 40 km/hr. and hav
ing a smooth steel surface. In order to prevent unsymmetrical wear of the
tread, the slip angle was varied in 30 seconds between +15° and -15°.
The cornering force versus slip angle was registered as record number 1.
MEASUREMENT OF TIRE PROPERTIES 713

After this first test the loaded tire was cycled 10 times between +10° and
-10°. Now, with a cycle time of 3 seconds, record number 2 was obtained
after 30 seconds [38].
The small 10 cycles were repeated, giving test number 3, and so on, but
in all tests the normal load and the inflation pressure were kept constant.
When after a number of tests reproducible results were obtained, with a
more or less equilibrium temperature of the tire, say after six tests, the av
erage value of the cornering force for the sixth to tenth test was considered
to be representative for the test tire in question, as in figure 8.3.57a. From
the above description it follows that to obtain comparable results for both
laboratory drum tests or road tests, it is necessary to "run in" test tires and
to control the temperatures and pressures. A common procedure, before
drum test, road trailer tests, or vehicle tests are executed, is that all tires
are run for break-in for a distance of about 300 km, at a speed of 80 km/
hr. and with no hard cornering maneuvers.
Limitations of test equipment
Although laboratory machines generally operate under controlled con
ditions, it is important to note, that due to effects of test speed and surface
curvature, force and moment data from curved surfaces cannot be conve
niently converted to comparable flat surface data. For example figure
8.3. 57b was plotted by dividing the lateral force (or aligning torque) values
from a 67 inch drum by the comparable value from the flat belt, showing
the significant effect of load at low slip angles. Other interesting results ob
tained are similar data for several machine combinations as shown in fig
ure 8.3.57c. Other limitations of test equipment are discussed in the litera
ture [47].

FIGURE 8.3.56. The conditioning of the tire.


714 MECHANICS OF PNEUMATIC TIRES

FIGURE 8.3. 57a. The differences in corneringforce-slip angle relationships of drum and road
tests.

GR70-I5 STEEL BELTED RADIALS- DESIGN "A"


28 PSI INFLATION PRESSURE
RATIOS OF MEAN VALUES OF 5-TIRE DATA SETS
gl.2

i...
LU
Sl.O
S
_,0.9

uj 0.8

* 0.7

0 I 2 16
I I I
25% Load Results are Based on
— VerySmoll Volues

I 0 I 2 4 8
a-SUP ANGLE, degrees

FIGURE 8.3.57b. Calspan 67 in roadwheel/Calspan TIRFflat belt-slip angle data


MEASUREMENT OF TIRE PROPERTIES 715

-GR 70-15 Steel Belted Radial-Design "A"


- 28 PSIG Inflation Pressure
- Ratios of Mean Values of S-Tire Sets
- Data Corrected for Speed Differences between
Machines
Colspon B«lt/BFG 1 20 Roodwheel
1.1

Calspon67/Calspon
Roodwheel/ Belt

BFG 120 Roodwheel /BFG Flo! Surface


09

124 8 12 16 20
a - SLIP ANGLE (DEGREES)

FIGURE 8.3.57c. Lateral force ratios at 100% of 24 pag (165 kPa) rated load

Mobile Tire Tests Versus Drum Tests


Due to significant differences induced by the curvature of the drum, the
cornering force coefficient and self-aligning moment coefficient obtained
from drum tests can be about (10-20%) lower than the values resulting
from flat dry concrete road tests. To illustrate. these differences in charac
teristics, the solid lines drawn in figure 8.3.57a show drum test results,
whereas the dashed lines give road test results obtained with the same
"run in" 165-15 tire. The self-aligning torque has been shown in figure
8.3.58. The drum diameter was 2.5 m.
To compare results of normal and experimental tires tested according to
the program described above under nominal constant load conditions, and
constant inflation pressure of 1.50 kgf/cm.2 at a speed of 40 km/hr., table
8.3.3 gives some results. The influence of the tire architecture is clearly
shown.
As discussed in chapter 8.2.5 the high speed tire tester of the Delft Ve
hicle Research Laboratory may be a powerful tool to obtain vehicle ori
ented tire data. In actual driving situations, parameters such as slip angle,
load and other variables may change rapidly during one tire revolution
and therefore cause deviations of tire properties obtained from steady
state tests. Transient tire properties obtained at realistic highway speeds
on different road surfaces will be meaningful, because mathematical mod
els are generally semi-empirical relationships using coefficients whose val
716 MECHANICS OF PNEUMATIC TIRES

16$ -IS , -Ms S>lf«l.gning torqu.


F,.«Skgf (
p, .150 kgf/cm? kgfm >i—1
^X
V .10 km/h t i/
r \
RiMd
Drum *
j \
2 J

[_ •
0 2 « 1 |

FIGURE 8.3.58. Illustration of the self aligning torque-slip angle relationships obtained on
drum and road tests.

ues are determined from tire test data [47]. It is not surprising that tire dy
namic properties have been given increased attention due to vehicle
handling in emergency maneuvers. However, the measurements of time-
dependent tire phenomena are difficult and require test facilities with
large range dynamic capabilities [40].
Mobile Tire Tests Versus Vehicle Tests on the Road
Having discussed drum tests versus mobile tire tests, it may be of inter
est to establish whether tire characteristics obtained from towed trailer
tests correlate with conventional vehicle tests on the road. The problem
encountered in the construction of a tire and the construction of the ve
hicle suspension is, what are the actual operating conditions. It is therefore
essential to establish whether certain trailer and laboratory tests are realis
tic or not. The tire characteristics obtained from the trailer tests as shown
in figures 8.3.25-28 already account for road surface irregularities, but it
still remains a problem to decide which tire out of a number of test tires is
the best from the standpoint of safety and vehicle handling properties. The
essence of the problem for the tire engineer is to know, what is the best
compromise in tire construction regarding tread, carcass stiffness, tread
compound, etc., in order to meet as far as possible the conditions given by
the vehicle suspension engineer.
Vehicle response tests determine vehicle handling properties by mea
surement of vehicle handling behavior in steady state cornering and tran
sient maneuvers, and by measurement of control modulation [39]. The ef
fect of the driver on vehicle behavior has been practically eliminated in
these tests.
To correlate mobile tire test results with vehicle response tests for a
given set of tires, it is necessary to determine the actual conditions encoun
tered on different road surfaces. Therefore, measurements of vehicle re
sponse tests should include the measurements of the tire forces on all
wheels, as well as the measurements of the path of the center of gravity of
the vehicle and the path of its individual wheels.
The measurement of the individual tire forces can be conducted with
MEASUREMENT OF TIRE PROPERTIES 717

8
- 1 + 1 + - 1 - T 1
-g 1* 1 + + +
D
|
3 • — —• *fi —• *c O O
1
o
e J 05

M i e
a z
f a
_c
8e
.!' r~ — CN — mot~ot~ —
— <S(S — —i<N—i — — IS
13 I I I I I I I I I I
g *
0
1 O

^ "S •^Vr^mV^vOinw^vO
3? e§
1

| ^J-^ilNmrorKtrl^t^t^t

g1°*
**i wi^or--oor-
00 1 1 - 1 1 1 + 1 1

2
H
o
O
II
5
13
«t
£
?
* a
g 2
i Q

£ 8
— so oo oo
— —
1 o
|*
O
1 1 TT
fS
1 •o
«
i?
i
i
<N — ONOOOOOO'TOv — —'

a
s
Q

P!
f 1
H - g-*
718 MECHANICS OF PNEUMATIC TIRES

the aid of the three-component road platform tester, as described in sec


tion 8.2.1 and shown in figure 8.2.17. The procedure of measuring the ac
tual tire forces with a road platform having the same surface as the test
track may be preferred over the method of using force transducers built
into each individual axle. It avoids the necessity of designing another axle
force measuring device for each vehicle type.
Two road platforms already provide much information, but if required
four of these platforms could be used. At present the author has only con
ducted experiments with one road platform, and load transfer could be
clearly demonstrated.
The measurement of the actual slip and camber angles when passing
over the road platform can be carried out with the S.C. meter as described
in section 8.3.3 and shown in figure 8.3.43.
The path over the- platform can be found by using the method of
mounting flashlights on the axle stub as shown in figure 8.3.42 and placing
a photo apparatus in a steel structure well above the road. The dashed
JLines obtained on the photograph indicate the path over the road platform.
"' The path of the center of gravity of the vehicle can be traced by filming
or photographing from a high tower made of a steel structure, placing
flashlamps on the front and rear ends of the test vehicle.
Of course, if required, the well known techniques of using accurate ac-
celerometers and gyros can be added to the instrumentation, but correc
tions for vehicle pitch and yaw motions have to made.
The disadvantage of the combination of these methods is the com
plexity of the total instrumentation system. However, it is for obvious rea
sons of great importance to relate mobile tire test data to those obtained
from vehicle performance.

References
This bibliography refers frequently to V.R.L.D. reports made by the Vehicle Research
Laboratory at Delft. These reports are in the Dutch language and can be obtained in micro
film or blueprint on payment.
[1] Freudenstein, G., Luftreifen bei Schräg- und Kurvenlauf, Deutsche Kraftfahrtfors-
chung 152 (1961).
[2] Grotewohl, A., Seitenwinduntersuchungen an Personenwagen, A.T.Z., 11/12 (1967)
[3] Vehicle Dynamics Terminology, Handbook Supplement, SAE J 670d (1975).
[4] Nordeen, D. I,., Analysis of tire lateral forces and interpretation of experimental tire
data, SAE Paper No. 670173 (1967).
[5] Bundorf, R. I, The influence of vehicle design parameters on characteristic speed and
understeer, SAE Paper No. 670078 (1967).
[6] Bergman, W., The basic nature of vehicle understeer-oversteer, SAE Paper No. 957B
(1965).
[7] Sabey, B. E., and Lupton, G. N., Photography of the real contact area of tires during
motion, Road Research Laboratory Report LR64 (1967).
[8] Cough, V. E., Practical tire research, SAE Trans. 64 (1956); Tire to ground stresses.
Wear 2(2) (Nov. 1958).
[9] Rooney, J. H. M., Measurements of the longitudinal deformation of a tread element,
V.R.L.D. Report No. P090 (1967).
[10] Iritani, S., and Baba, 1 ., Forces on the contact patch of the tire, Paper B6, Proc. 10th
FISITA Congress, Tokyo, 1964.
[1 1] Gough, V. E., Nondestructive estimation of resistance of tire construction of tread wear,
SAE Paper No. 667A (1963).
[12] Gorp, H. A. van. Determination of lateral accelerations acceptable by motorists,
V.R.L.D. Report No. 369 (1966).
[13] Dijks, A., Timan, D. A., and Ruyter, T. J., Trailer road tests versus vehicle tests on a
smooth polished road surface, V.R.L.D. Report No. PI 19 (June 1969).
MEASUREMENT OF TIRE PROPERTIES 719

[14] Holmes, K. E, and Stone, R. L>., Tyre forces as functions of cornering and braking slip
on wet road surfaces, Road Research Laboratory Report LR254 (1969).
[15] Buis, P., Pneumatic tires, V.R L.D. Report No. 084 (1967).
[16] Koeszler, P., and Senger, G., Vergleichende Untersuchungen der Seitenfilhrungseigens-
chaften von Personenwagen Reiten, Deutsche Kraftfahrtforschung 172 (1964).
[17] Henker, E., Dynamische kennlinien von P.K.W. reifen (V.E.B. Druckerei, D.D.R.);
Wissenschaftlich-technische Veröffentlichungen aus dem Automobilbau. H3 (1968).
[18] Allbert, B. J., and Walker, J. C, Tyre to wet road friction at high speeds, Proc. IME
180, 105 (1965/66).
[19] K on ing, C. J. de, Design considerations of a device for measuring the camber- and slip
angle, V.R.L.D. Reports No. 398a,b,c (1967).
[20] Lems, F. L... The measurement of camber and slip angles of motorcars, V.R.L.D. Re
ports No. 464a,b (1968).
[21] Cues, C. G., Lander, F. T. W., and Holmes, K. E., A test vehicle for studying the skid
resisting properties of tyres and road surfaces under controlled cornering and brak
ing. Road Research Laboratory Report LR99.
[22] Davisson, J. A., Basic test methods for evaluating tire traction, SAE Paper No. 680136
(Jan. 1968).
[23] Kummer, H. W., and Meyer, W. E., Tentative skid-resistance requirements for main ru
ral highways, Nat. Coop. Highway Res. Program Report 37 (Highway Research
Board, Washington, D.C. 20418).
[24] Savkoor, A. E., The isothermal friction of rubber compounds on nominally smooth
hard tracks, V.R.L.D. Report No. P093 (1968).
[25] Spaink, G. N., Friction and drainage measurements on abrasion paper of different qual
ities, V.R.L.D. Report No. 391 (Apr. 1967) (Translated by Cornell Aeronautical Lab
oratory, Inc., Buffalo).
[26] Schulze, K. H., and Beck man n. L... Friction properties of pavements at different speeds,
A.S.T.M. Spec. Tech. Publ, No. 326 (1962).
[27] Sabey, B. E., Wet road skidding resistance at high speeds on a variety of surfaces on A
1, Road Research Laboratory Report LR131 (1968).
[28a] Moore, D. F., Drainage criteria for runway surface roughness, Cornell Aeronautical
Laboratory, Buffalo, J. Roy. Aero. Soc. (London) 69 (653) (1965).
[28b] Moore, D. F., A history of research on surface texture effects, Wear 13, 381-412
(1969).
[29] Edelman, A., Measurements with a refined drainage meter, V.R.L.D. Report No. 473
(Sept. 1968).
[30] Ernst, H. C., Correlation of results of measurements of coefficients of friction obtained
with a road trailer tire tester, a skid resistance tester (S.R.T.) and a modified Dr.
Moore drainage meter, V.R.L.D. Report No. 4971 (June 1969).
[31] Ernst, H. C., Development of a Delft high pressure drainage meter, V.R.L.D. Report
No. 497" (Aug. 1969).
[32] Grosch, K. A., and Maycock, G., Influence of test conditions on wet skid resistance of
tyre tread compounds, Trans. I.R.I. 42(6) (1%6).
[33] Gillespie, T. D., Pavement surface characteristics and their correlation with skid resis
tance, Perm. Dept. Highways Program Report No. 12 (1965).
[34] Gerritsen, R. R. V., Description of portable pavement friction testers. Design of a pen
dulum tester, V.R.L.D. Reports No. 392-395 (1967).
[35] Koelewijn, A. C., Pavement friction testers, V.R.L.D. Report No. 516 (1969).
[36] Paar, H. G., Skid trailer for testing truck tires, V.R.L.D. Report No. 376 (Nov. 1966).
[37] Schrier, J., and Groeneweg, H. H., Design and calculation of a skid trailer for testing
truck tires, V.R.L.D. Report No. M077 (May 1968).
[38] Timan. D. A., and Ruyter, T., Cornering force and self-aligning torque measurements
on a drum and dry road with the tire tester, V.R.L.D. Report No. P100 (Oct. 1968).
[39] Bergman, W., Considerations in determining vehicle handling requirements, SAE Pa
per No. 690234 (Jan. 1969).
[40] Scouring, D. J., "Dynamic Response of Tires," Tire Science and Technology, TSTCA,
Vol. 4, No. 2, May 1976, pp 1 15-145.
[41] Dijks, A., "A Multifactor Examination of Wet Skid Resistance of Car Tires," S.A.E.
Paper 741 106, October 1974.
[42] Peterson, K. G., Smithson, F. D., and Hill, F. W., "General Motors Tire Performance
Criteria (TPC) Specification System, S.A.E. Paper 741103, October 1974.
[43] Tielking, J. T., Fancher, P. S., and Wild, R. E., "Mechanical Properties of Truck Tires,"
S.A.E. Paper 730183, January 1973.
[44] Savkor, A. R., "Adhesion and Deformation Friction of Polymer on Hard Solids," Ad
vances in Polymer Friction and Wear, Vol. 5A, 1974, Edited by Lieng-Huang Lee,
Plenum Publishing Corporation, New York and London.
720 MECHANICS OF PNEUMATIC TIRES

[45] Dijks, A., "Wet Skid Resistance of Car and Truck Tires," Tire Science and Technology,
TSTCA, Vol. 1, No. 2, May 1974, pp. 102-116.
[46] Pottmger, M. G., Marshall, K. D., and Arnold, G. A., "Effects of Test Speed and Sur
face Curvature on Cornering Properties of Tires, S.A.E. Paper 760029, February
1976.
[47] Bergman, W. and Clemen, H. R , 'Tire Cornering Properties," Tire Science and Tech
nology, Vol. 3, No. 3, Aug. 1975, pp. 135-163.
[48] Moore, D. F., "Principles and Applications of Tribology," Pergamon Press, Oxford-
New York, Edition 1975.
[49] Moore, D. F., "The Friction of Pneumatic Tyres," Elsevier Scientific Publishing Com
pany, Amsterdam, 1975.
Chapter 9
ANALYSIS OF TIRE PROPERTIES
H. Pacejka1

CHAPTER 9

9. 1Introduction 722
9.2 Nomenclature 723
9.3 List of Symbols 723
9.4 Tire In-Plane Dynamics «. 726
9.4.1 Low frequency properties 726
9.4.2 High frequency properties 757
9.5 Yaw and Camber Analysis 785
9.5.1 Steady State Motion 790
9.5.2 Nonsteady State Motions 833

1 Laboratory for Vehicle Dynam.cs. Delft Technological University, Delft. Netherlands.

721
722 MECHANICS OF PNEUMATIC TIRES

9.1 Introduction
The combination of road, tire, vehicle, and driver forms one entity. The
mechanical characteristics of the tire in contact with the road must com
bine with the mechanics of the vehicle to help in producing operational
characteristics of the tire-vehicle system which are satisfactory to the
driver.
"The complexity of the structure and behavior of the tire are such that
no complete and satisfactory theory has yet been propounded. The char
acteristics of the tire still present a challenge to the natural philosopher to
devise a theory which shall coordinate the vast mass of empirical data and
give some guidance to the manufacturer and user. This is an inviting field
for the application of mathematics to the physical world."
In this way Temple formulated the situation of more than one decade
ago (Endeavour, October 1956). Since then, in numerous institutes and
laboratories, the work of the earlier investigators has been continued.
Considerable progress in the development of tire mechanics during the
last decade has led to a better understanding of tire behavior. Owing to the
infinite complexity of the pneumatic tire and its interaction with the road
it does not appear at present, despite the progress made, that Temple's
view will be altered in the foreseeable future. Thanks to new and more re
fined experimental techniques becoming increasingly available, and to the
introduction of the electronic computer, the goal of formulating more real
istic mathematical models based on better insight and leading to more re
liable prediction of tire performance may be achieved.
The author of this chapter does not claim to have supplied a picture of
tire behavior which covers all knowledge achieved hitherto. A selection of
studies has been made in order to provide the engineer and the student
with background material necessary for the investigation and the under
standing of tire and vehicle functional performance.
From the point of view of the engineer and the applied mathematician
the mechanical behavior of the tire must be systematically investigated in
terms of its reaction to various kinds of input related to vehicle motions
and road parameters.
With reference to the role of a tire it is convenient to distinguish be
tween symmetric and anti-symmetric modes of performance. First, the tire
supports the vertical axle load and transmits longitudinal braking or driv
ing forces. Second, the tire is called upon to supply the lateral cornering
and camber forces which are necessary for the directional control of the
vehicle.
The content of this chapter has been subdivided according to these cate
gories. In addition to steady-state or slowly varying motions also high-fre
quency and n onsteady state behavior of the tire have been treated. Experi
mental results have also been added.
Many of the investigations discussed in this chapter have been carried
out at the Vehicle Research Laboratory of the University of Technology,
Delft, Holland.
The author wishes to express his appreciation to the members of the
staff of this laboratory: Especially to A. Dijks for contributing to parts of
the text as well as reviewing parts during its preparation, J. van den Berg
and J. A. Zwaan; E. G. M. J. de Vries and his electronic measuring depart
ment; D. A. Timan, J. H. M. Rooney and P. J. Jillesma for their numerous
tire experiments, as well as to H. M. Snijders and his workshop for the as
sistance in manufacturing various instruments and apparatus.
ANALYSIS OF TIRE PROPERTIES 723

9.2 Nomenclature
For both the experimental and theoretical investigations of tire behavior
described in this chapter, we have attempted to use a uniform system of
notation. As a rule, the meaning of symbols has been explained in the text.
For this reason, only a list of the most important symbols will be given be
low. The choice of symbols has been inspired by the list which has been
proposed by a SAE committee last revised in 1974 (Vehicle Dynamics
Terminology SAE J670d, published July 1975 by SAE). A number of
changes and additions appeared to be necessary in order to obtain a more
or less systematic and usable system of symbols adjusted to the specific
subjects of this chapter.
Constant quantities describing construction, configuration and proper
ties of the real tire or of the theoretical model are defined in such a way
that they become positive. In most cases, the positive sense of variable
quantities are chosen in accordance with the (C, x, y, z) system of axes
shown in figure 9.2.1. The origin C, defined as contact center, is the point
of intersection of the road-plane, the wheel center-plane and the plane
which is situated normal to the road-plane and which passes the wheel
axis. The x-axis points forward and forms the intersection of the wheel
center-plane and road-plane. The z-axis points downward and is directed
perpendicular to the road-plane. Consequently, the y-axis is the per
pendicular projection of the wheel axis onto the road. In the same figure
9.2.1, the positive directions of forces and moments acting from road to
tire have been indicated, as well as the positive senses of the variables
which describe the deviations of the position and the motion of the wheel
center-plane with respect to the rectilinear steady state motion of the
wheel center-plane, which in that case coincides with the (x, £) plane of
the coordinate system (0, x, y, z) fixed to the road with the z-axis directed
vertically. As in figure 9.2.1, the road-plane has in most cases been consid
ered as a smooth horizontal surface.
In some cases an alternative definition of positive sense has been felt to
be preferable. In order to work with positive quantities, the tire normal
load has been defined as FN(=W) = -FZ. Similarly, the quantity Fr =
-Fx has been introduced, denoting the rolling resistance force during free
rolling, i.e., at constant forward velocity and without traction or braking
torques. Also, the sense of the speed of rotation SI of the forward rolling
wheel has been defined as positive. Sometimes, the absolute values of the
longitudinal force Fx have been considered. They are designated as the
braking force FJ= —Fx) and the traction force FT(=Fx).
The lateral force acting from road to tire, F^ has been provided with an
additional subscript a or γ in cases when it has been felt necessary to ex
press whether side slip or camber causes the lateral force.

93 List of symbols

L, F and T denote length, force and time units respectively.
RAD denotes radians.
a half length of contact area (L)
b half width of contact area (L)
cw, foundation stiffness per unit length in tangential (/), lateral (c)
and radial direction (r) respectively (F/L2)
724 MECHANICS OF PNEUMATIC TIRES

normal to road-plane

wheel-centre-plane

wheel-axis

steady state (stationary) general deviated


rectilinear free rolling. situation.

FIGURE 9.2. 1 . Nomenclature and coordinate system for a wheel on a plane surface.

cr,j, stiffness of tread rubber per unit area in longitudinal (x) and
lateral direction (y) respectively (F/L3)
c, tensile tread band (carcass) stiffness per unit length (F)
Cfcf carcass stiffness in contact region in x and y directions respec
tively (F/L)
CF, (=dFy/da at a=0) cornering stiffness (cornering rate) (F/RAD)
CFr (=dFf/dy at y=a=0) camber rate (cf. eq (9.5.61)) (F)
CMa (=-dM,/da at a=0) cornering stiffness (aligning rate) (FL/
RAD)
C, (=3Mr/df at 5=60) rolling resistance coefficient (F)
C. (=-dFJdxa or BFJdr,, at F,=0) longitudinal or tangential
stiffness in contact region of non-rolling tire (F/L)
Cy (=—dFJdya at F=0) lateral tire stiffness in contact region of
non-rolling tire (F/L)
C, (=3W/df=dW/dza at S=S0 or za=0) normal tire stiffness (F/L)
C. (=dF,/dK at ic=0) longitudinal slip stiffness (F)
C+ (=—dMJd^> at A/,=0) torsional stiffness about vertical axis of
non-rolling tire (FL/RAD)
El flexural rigidity of tread band (FL2)
Fa.r braking and traction force respectively (F)
FN (-W—F,) tire normal load (F)
ANALYSIS OF TIRE PROPERTIES 725

F,^ longitudinal, cornering (^lateral) and normal force acting


from road to tire (cf. fig. 9.2.1) (F)
G{x, £) Green's function
7X polar moment of inertia of wheel about wheel axle (FLT2)
I wavelength of standing wave (L); half of projected contact
length (cf. fig. 9.5.1 of sec. 9.5)
M, (=Mr) rolling resistance moment (cf. fig. 9.2.1.) (FL)
M,yJ moment acting from road to tire (cf. fig. 9.2.1), M, — aligning
torque (FL)
n frequency of motion (Hz)
P*#j contact force per unit area acting upon tire in negative x, y and
z direction respectively (F/L2)
p, inflation pressure (F/L2)
<7W contact force per unit length acting upon tire in negative x, y
and z direction respectively (F/L)
r tire radius (L)
r. effective radius of rolling (L)
r, loaded tire radius = wheel center height (L)
R radius of curvature (L)
s distance travelled (L); mode number
S tension force in tread band (F)
/ time; pneumatic trail (=—M[/F1,) (L)
u tangential (longitudinal) deflection (L)
v lateral deflection (L)
vc . lateral deflection of carcass and tread rubber respectively (L)
V (=ds/dt) speed of travel (L/T)
Vw creep (slip), rolling and sliding velocity respectively (L/T)
H- radial deflection (positive outwards) (L)
W (=-F.) tire normal load (F)
X longitudinal horizontal force acting upon tire (F)
x,y,z coordinates with respect to moving system (fig. 9.5.1)
x,y,z coordinates of contact center C with respect to system fixed in
space
*«> y*, z<, variation of wheel center position with respect to steady state
motion
xa ya zc coordinates of contact point with respect to system fixed in
space
a slip angle (cf. fig. 8.1.1); crown angle
P (=arctan dy/ds) path angle (cf. fig. 8.1.1)
v camber angle of wheel center plane (cf. fig. 8.1.1)
8 normal tire deflection (positive toward the center) (L)
X wavelength of motion (L)
K longitudinal slip value (cf. eqs (9.4.27, 9.5.64)
Ka r percentage of brake and traction slip respectively
ft coefficient of friction
<f> (=tty/ds) spin = yaw rate (RAD/L)
if/ yaw angle (cf. fig. 8.1.1)
X deviation from steady state angle of rotation Qt (cf. fig. 9.4. 13)
p mass density of tire tread band (FT2/L2)
a relaxation length for tire model without tread rubber (L)
a* relaxation length for tire model with tread rubber (L)
co frequency of motion (RAD/T)
co, reduced, spatial or path frequency (RAD/L)
Q speed of rotation of wheel (RAD/T)
726 MECHANICS OF PNEUMATIC TIRES

9.4. Tire In-Plane Dynamics


In this part the symmetric behavior of tires will be discussed. Section
9.4. 1 is devoted to relatively low-frequency phenomena with inertia effects
of the tire neglected. Attention will be paid to the normal and longitudinal
problem. Section 9.4.2 treats the high-frequency vibrational behavior of
tires,a gain mainly restricted to symmetric or in-plane aspects.
The frequency below which tire inertia effects may be neglected must lie
well below the lowest natural frequency of the tire. Distinction may be
drawn here between radial and longitudinal motions and between differ
ent types of tire construction. However, it is estimated that in general the
low frequency range can be extended up to about 20 Hz. It may be noted
that for lateral tire movements this limit must be reduced (cf. sec. 9.5.3 on
gyroscopic effects).

9.4.1. Low Frequency Properties


Normal Force-Deflection Characteristics
It has been generally accepted that pneumatic tires transmit their nor
mal load mainly by the formation of a finite flat contact area, A, which en
ables the internal air pressure, ph to remain in equilibrium with the ex
ternal vertical contact pressure, p,. We would therefore expect the normal
load, W, to be approximately equal to the inflation pressure multiplied by
the contact area. This would indeed apply for a thin envelope or mem
brane. However, the tire tread-band in particular cannot be considered to
be thin. Automobile tires often show nearly rectangular foot prints. This is
due to the nearly flat profile of the undeflected tire cross section. The ef
fective area, defined as Ae = W/pn then becomes less than the gross foot
print area A. This is because the contour of the presumably thin side wall
cross section does not show a horizontal tangent at the point where the
contact area begins. The tension force in the side wall, caused by the air
pressure, has an upward component which reduces the resulting pneu
matic force. Later on in this paragraph an analysis of this phenomenon
will be given. Use will be made of a tire model showing thin flexible side
walls and a rigid tread-band cross section. In this anlaysis the internal
pressure pi has been kept constant. For automobile tires this is an accept
able assumption. The effect of pressure rise may not be negligible for air
craft tires because of their relatively large deflections [1].1
In the absence of internal air pressure the tire model proposed is not ca
pable of transmitting a radial force if one neglects the flexural rigidity of
the tread-band. Consequently, the vertical stiffness of the tire model is of a
completely pneumatic nature. In reality.however, this is not entirely so.
Experiments of various investigators [1,2, 13] indicate that the rigidity of
the cover (side walls and tread-band) causes a noticeable contribution to
the force transmission. According to these references, under rated inflation
pressure conditions the carcass carries about 15 percent of the vertical load
for cross ply automobile tires and 3 percent to 8 percent in case of aircraft
tires. As a result of this cover rigidity the effective area Ae increases and
may even exceed the gross foot print area A.
1 Figures in brackets indicate the literature references at the end of this lection.
ANALYSIS OF TIRE PROPERTIES 727

The above observations hold for standing or nonrotating tires pressed


against a flat surface. On curved surfaces the tension force in the tread-
band will take part in the force tranmission. In addition, we observe that
with given radial deflection at the contact center, the contact length
changes with curvature of both tire and surface. It is therefore to be ex
pected that experiments indicate a dependence of normal tire stiffness
upon road surface curvature. Both Marquard [3] and Chiesa [4] found a
decrease in stiffness with decreasing radius of curvature of the surface.
Marquard has shown that the influence of surface curvature can be
roughly approximated with the following equation:

(9.4.1)
R+r

In this equation, which originates from the Herzian theory of two parallel
cylinders pressed against each other, C , denotes the normal tire stiffness
expressed as tire force per unit deflection, R the radius of the contact sur
face and r the tire radius. Marquard found the formula to follow the ex
perimental trend for ratios as small as R/r = 0.25. When traversing obsta
cles with radii of curvature much smaller than the tire contact length, the
tire clearly demonstrates its more than zero-dimensional nature. A sub
sequent section treats the enveloping properties of tires. The remainder of
this section will be confined to the contact with flat surfaces.
Until now, the tire has been considered as nonrotating. Once the tire
rolls, fresh elements of cover are continuously entering the deformation
region. There is no a priori reason to believe that deformation of a rolling
tire follows the same rules which hold for a standing tire. Hysteresis,
which has been found to damp the vertical motion of the axle with a non-
rotating wheel, appears to be practically absent with a rotating tire once
the rate of rolling becomes high as compared to the rate of deflection. In
stead, hysteresis produces rolling resistance. In addition, it appears that
vertical tire stiffness is affected by the rolling process. Rasmussen and Cor-
tese [5] determined the effective tire stiffness by means of resonant tests.
They show that the effective normal stiffness of a rolling tire is virtually
independent of hub amplitude whereas the effective stiffness of a non-
rolling tire varies with amplitude in a nonlinear fashion. At small ampli
tude the standing tire shows considerably larger values for the stiffness
than the rolling tire (50% higher at 1.5 mm. amplitude for a 6.50-14 tire).
For increasing amplitudes the stiffness of the standing tire decays gradu
ally and tends to the constant value of the rolling tire (15% higher at 10
mm.). Chiesa and Tangorra [4] found with their resonance tests that this
level of stiffness is not very much affected by the speed of rolling once the
speed has exceeded a value of about 20 km/hr. (cf. fig. 8.2.28a). The above
observations are of particular importance for the execution of laboratory
ride simulator tests (cf. Betz [6]).
At high rolling speeds the dynamic aspect, which among other things is
responsible for the formation of standing waves (sec. 9.4.2), must be taken
into account. Dodge [7] made an attempt to attack this problem. With the
aid of equations describing the dynamics of a rotating shell, corresponding
to eqs. (8.2.90-91), the radial stiffness of the shell subjected to a radial
point load has been determined. The results are complicated and difficult
to interpret and are not directly applicable to the tire pressed upon a flat
surface. An extension of the analysis is needed where the following addi
728 MECHANICS OF PNEUMATIC TIRES

tional parameters are introduced: finite contact length, internal pressure,


tension force in tread-band and tangential stiffness of tread-band with re
spect to the wheel rim. Another theory related to the work of Dodge has
been proposed by Clark [43]. In this study the contact length has been
considered as finite.
Before going into details of the static loading process, we consider the
following simple but interesting observation. For small normal tire deflec
tions the contact length varies approximately as the square root of the de
flection. Consequently, for a foot print of constant shape the contact area
varies as the deflection. This means that, with the assumption that the nor
mal load equals internal pressure multiplied by a fixed percentage of the
contact area, the force-deflection curve is linear for small values of the de
flection.
For a more sophisticated analysis of the normal force-deflection prob
lem we make use of the membrane concept employed by Rotta [8] and
Senger [9] among others for the analysis of the elastic properties of a tire
segment, and by Pacejka [10] for the investigation of air springs. Senger
used the model described earlier, consisting of flexible side walls and a
tread-band with rigid cross section. Senger replaced the circular band by a
linear elastically-supported beam under tension. The beam is transversely
deflected by the application of a normal force distributed over a finite
length. The longitudinal tension force influences the resulting transverse
stiffness due to the curvature caused by the application of the normal
forces. In the writers opinion the transformation to a linear beam is not
valid for analyses of radial deformations. In reality, the contact area is not
curved when pressed on a flat road. Therefore the tension force, if present
in this area, cannot influence the force distribution in the contact area.
Flexural rigidity of tread band and side walls produce additional normal
forces but will be neglected henceforth.
Theoretical load-deflection relationship
For the determination of the vertical force acting on a deflected tire we
divide the tire into a large number of thin segments, and imagine this force
to be composed of the elementary radial forces acting on these segments.
For the calculation of the elementary forces we need to know the load-de
flection characteristic of one segment. This characteristic will be desig
nated as the radial foundation characteristic.
We consider the model of the tire cross section shown in figure 9.4. 1 .
The side wall is assumed to behave as an inextensible membrane. For the
segment having a nearly uniform thickness, the membrane assumes a cir
cular shape under the action of the internal air pressure pi. The stresses in
the membrane would not change if the membrane were extended accord
ing to the dashed line shown in the figure. Evidently, the external vertical
force acting on the segment of unit thickness is

fc -/*-** (9-4-2)

with 2be denoting the effective width indicated in figure 9.4.1. With tire
parameters l and /*„, introduced in the figure, we obtain the following
equations for b, and the radial deflection — w in terms of the parameter <t>,

(9.4.3)
ANALYSIS OF TIRE PROPERTIES 729

(9.4.4)

The arc length / is assumed to be a given constant. At zero deflection, i.e.,


w = 0, the effective side wall height is hm. Due to the tensile rigidity of the
tread band, the effective width be does not, in general, vanish in the unde
tected case. The resulting radial force produces the hoop tension force
which is present in the undeflected portion of the tire tread band.
These observations indicate that this characteristic will in general not
pass through the origin (q, = 0, w = 0). The foundation characteristic can
be calculated from the equations above. For a segment with relative di
mensions l = 1.75 b, h,n— 1.6 b, the dimensionless foundation character
istic is presented in figure 9.4.2. From the equations (9.4.3-4) the stiffness
of a segment in the radial direction (radial foundation stiffness per unit
length) can be derived. We obtain:

-2pt dw
db,\ cos <j>, + & sin <j>,
(9.4.5)
sin <f>, - <J>, cos <t>,
Together with eq (9.4.3) cr can be calculated as a function of w.
In addition, we derive the lateral stiffness cc of the tire segment. With a
small lateral displacement ν of the tread band in lateral direction y, point
A (fig. 9.4.1) moves upward and point B downward a distance (b-be)ν/hs.
The total lateral force acting on a segment of unit thickness becomes
2pKb-be)ν/hr Hence, the lateral stiffness of the segment (lateral founda
tion stiffness per unit length) reads:
b-b.
(9.4.6)

FIGURE 9.4. 1 . Real and model representation of tire cross-section.


730 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4.2. Radialfoundation characteristics with radial stiffness, cn and lateral stiffness,
ca of tire segment of unit thickness shown in figure 9.4.1.

The lateral stiffness in an appropriate dimensionless form appears to bear


a simple relationship to the dimensionless radial foundation stiffness, as
shown in figure 9.4.2. From this figure it can be observed that both the
radial and lateral stiffness decrease with increased deflection (-w). The
lateral stiffness even becomes negative for be > b. The lateral stiffness is of
particular importance in the analysis of the cornering behavior of tires, to
be dealt with in part 9.5. Rotta [8] and Clark [11] employed alternative
models for the determination of the lateral foundation stiffness.
We shall continue now with the normal loading problem and determine
the load-deflection curve of the tire from the foundation characteristic of
figure 9.4.2. The radial deflection —w varies along the contact line. The
contact length depends on the normal tire deflections, δ, and the tire
radius r. We use the approximate equations
cf - 2r8 (9.4.7)
and

(9.4.8)

where x denotes the coordinate in the longitudinal direction. In figure


9.4.3 a comparison is made between calculated values of contact length
and experimental data from Senger [9]. Note that at large deflections the
calculated values are somewhat high.
ANALYSIS OF TIRE PROPERTIES 731

The dimensionless normal load expressed in the integral form


W ^-d (9.4.9)

can be calculated, in principle, with the aid of the foregoing equations. In


the actual calculation the radial foundation characteristic of figure 9.4.2
has been approximated by a quadratic function of w/b, coinciding with
the original curve in -w/b = 0, 0.2, and 0.4. The final result for a tire with
r = 5.4 b has been shown in figure 9.4.4 in comparison with measured data
deduced from Senger's experimental results. Perhaps owing to deviations
between calculated and measured values of contact length (fig. 9.4.3), the
agreement is better than expected. As has been mentioned before, around
15 percent of the normal force transmission is attributable to carcass rigid
ity. The shape of the measured and calculated curves is representative for
cross-ply tires. It suggests that for most purposes a linear load-deflection
relationship can be used with sufficient accuracy. Vibration tests of Ras-
mussen and Cortese [5] and of Betz [6] with rolling tires substantiate this
result.
Tiemann [12] found great similarity between reduced load-deflection
curves for a great variety of automobile and truck tires. Here, distinction
must be made between bias-ply and radial tires. For each of these cate
gories of tires it turns out that the dimensionless-load versus dimensionless
deflection can be reasonably well represented by a single curve
W
(9.4.10)

where B is section width, H is section height and r is rim radius. For bias-
ply tires the doubly-curved form of figure 9.4.4 applies, whereas radial

234
Deflection 8 (cm >

FIGURE 9.4.3. Experimental and calculated values of contact length f9J.


732 MECHANICS OF PNEUMATIC TIRES

u
w

as
. calc u lated
measurtd

0.1 02 0.3 04

FIGURE 9.4.4. Calculated nondimensional load deflection curve compared with experimental
results ofSenger [9] for a 11.00-20 cross-ply truck tire (p, - 4-7 bar).

tires show a continuously concave characteristic. For aircraft tires Smiley


and Home [13] propose a similar equation. To account for the influence of
cover rigidity they replaced pi by pi + 0.08 pn with pr denoting the rated
inflation pressure. Since the measured curves are so similar to those given
by eq (9.4.10), it seems possible that vertical load-deflection curves could
be calculated for each type of tire construction.
Enveloping Properties (Obstacles)
An important property of a pneumatic tire is its ability to cushion a ve
hicle against short road irregularities. A rigid wheel passing over an ob
stacle would acquire a sudden vertical velocity which involves extremely
large vertical accelerations. Experiments never show such large accelera
tions so that obviously the elastic enveloping properties enable the pneu
matic tire to partially "swallow" the obstacle while rolling over it. The re
sulting vertical displacement of the axle is small relative to the height of
the obstacle.
Most of the published information on this subject concerns experimen
tal data relating tire parameters to the response of tire forces or wheel axle
motions as the tire slowly rolls over an obstacle. A number of investigators
have examined the response of the vertical and longitudinal tire forces to a
short prismatic obstacle extending over the whole tire width, and possess
ing a rectangular, trapezoidal, sinusoidal or cylindrical cross section. In
these tests the axle height above the road plane has been kept constant. As
an illustration we present in figure 9.4.5 (from Gough [14]) the response of
a bias-ply and a radial tire. It is seen that in the variation of the vertical
load W = -Fz between two maxima, a minimum arises when the wheel
axle is located directly above the center of the obstacle. In this particular
situation the minimum becomes even lower than the initial normal load
-Fzo occurring on a flat surface. Extensive tests of Julien [15] show that
the shape of the curves vary both with size and shape of the obstacle and
with the vertical deflection of the tire. It appears that the minimum in ver
tical force variation only arises when the static tire deflection is sufficiently
large and the obstacle size sufficiently small. The longitudinal force in
ANALYSIS OF TIRE PROPERTIES 733

variably shows one maximum and one minimum. These extreme values
are virtually independent of initial tire deflection [15].
From experiments conducted by Lippmann and his associates [16, 17] it
appears that with fixed axle height the vertical and longitudinal peak
forces vary nearly proportionally with internal pressure. For ordinary in
flation pressures the major part of the forces arise from internal pressure.
This leads Lippmann to the conclusion that the core of the process rests on
some pneumatic mechanism.
Figure 9.4.5 furthermore shows the wheel rotation per unit of travel on
the road. The angular acceleration of the wheel caused by obstacles will
generate additional longitudinal forces which may become quite consid
erable in magnitude, particularly at high speed.
Theory of enveloping capabilities
An adequate theory explaining these measured force variations has not
been found in the existing literature. The following relatively simple the
ory may furnish some insight into the problem. We employ a tire model
consisting of a large number of radially directed springs. The influence of
tread band tension and bending stiffness will be neglected. It has been
pointed out in section 9.4.1 that the spring forces are mainly due to pneu
matic action. It is assumed furthermore that these forces are directed per
pendicular to the tire peripheral line. For the sake of simplicity the shape
of the peripheral line is considered to remain circular outside the contact
zone, which extends from the leading to the trailing contact points. When,
in addition, the stiffness of the springs is constant i.e., linear foundation
characteristic, the vertical force which acts on the tire can be obtained ap
proximately by multiplying the overlapping area of the tire's circum
scribed circle and the road profile by the foundation stiffness per unit
length of circumference. According to the behavior of such a tire model,
the vertical force increases when a short obstacle is encountered. The force
remains constant as long as the obstacle contour lies completely inside the
circumscribed circle. The experimental evidence of the occurrence of a
minimum force cannot be explained with this simple model unless a non
linear softening foundation characteristic is assumed. As has been in
dicated in section 9.4.1, such a spring characteristic of a tire element will
indeed exist (fig. 9.4.2). During the time in which the obstacle is com
pletely swallowed by the tire, it is obvious that the vertical force becomes a
maximum when the obstacle is in the foremost and rearmost positions
since the slope of the load-deflection curve is then greatest. In the center
position a minimum is expected as the obstacle now deforms the tire in the
range of lowest stiffness.
Experiments indicate that under particular conditions a minimum verti
cal force can arise which is even lower than the initial force without an ob
stacle. Since the force-deflection characteristic of a tire element is not ex
pected to be particularly nonlinear in the practical range of deflection,
some other mechanism must be responsible for this phenomenon.
It is believed that a second possibly important effect on enveloping an
obstacle is the shrinkage of the circumference of the circumscribed circle
of the tread band, which is assumed inextensible in the ensuing analysis.
This circumference must become shorter in order to supply length in the
contact zone where the obstacle is partially surrounded by the tread band.
In figure 9.4.6 two circumscribed circles are shown. The larger circle
shows the tire pressed against a flat surface. Its size reduces to the smaller
734 MECHANICS OF PNEUMATIC TIRES

ilOO

Position of obstacle relative


I I I I'

Position of obstacle relative


1 1 I ff '° °"e

4 8 12 16
Distance travelled byroad, In.

FIGURE 9.4.5. Variation of normal load and dragforce as well as angular speed of wheel rim
as a function of obstacle position with respect to wheel axle [14],
ANALYSIS OF TIRE PROPERTIES 735

circle when the obstacle is in the contact center. It has been assumed that
the tread band deformations caused by the intrusion of the obstacle vanish
at the top of the tire. Consequently both circles touch each other in this
point. The difference in length of the two circumferences equals the differ
ence between obstacle contour length and its base length. The relatively
small variation of the difference between arc length and chord length
(contact length) of the circles in these two cases has been neglected. The
possibility of partial loss of contact has been disregarded, an assumption
which is admissable only in case of relatively smooth obstacles.
The reduction of the circumference of the circumscribed circle has been
designated as 2λh, in which h denotes the obstacle height and λ a non-
dimensional form parameter of the obstacle. For a rectangular obstacle
shape, λ = 1 . The value of λ decreases when the shape becomes trape
zoidal. The decrease in diameter of the circumscribed circle becomes 2λh/
it. For small values of h the area of the section with chord length 2a and
width 2XA/7T is approximately 4λha/ir. With cr denoting the radial (verti
cal) stiffness of the foundation per unit area, Fzo the initial vertical load
and A the area of the obstacle cross section, we obtain for the vertical load
acting on the tire when the obstacle has arrived directly below the wheel
axle:
4\ha
-F. = -F,, + c. A - (9.4.11)

It is seen that from this formula a decrease in vertical force can indeed oc
cur even using linear foundation characteristics. The condition at which
this occurs is:
4\ha
A< (9.4.12)

original
reduced
circumsc ri bed
circle of tyre
peripheral line

FIGURE 9.4.6. Tire peripheral line with and without the intrusion of an obstacle.
736 MECHANICS OF PNEUMATIC TIRES

It is of importance to introduce another form factor a defined by the rela


tion:
A = ah2. (9.4.13)
For a square cross section σ = 1 . When the length becomes larger than the
height a increases. The condition (9.4.12) reads in dimensionless form:
A<1A (9.4.14)

Julien [15] has carried out experiments with an obstacle of square cross
section. We wish to compare his experimental results with our theory. For
this purpose we adopt a trapezoidal obstacle to the shape of which the real
tire deforms reasonably closely when it rolls over a square obstacle. The
symmetric trapezoid chosen has a height h, a base line 3h and a top line h.
The area becomes A = 2h2 so that σ = 2 and the parameter \ = J2 - 1 =
0.414. The condition (9.4.14) then becomes h/a < 0.263. From the results
of Julien's experiments with a 5.0-15 tire it can be deduced that the mini
mum vertical force becomes less than the initial load at the same axle
height when the ratio obstacle height to half contact length, h/a, becomes
less than the values shown in the table below, valid for three values of ini
tial tire deflection δ.
The agreement with the theoretical value 0.263 is very good considering
the simplicity of the model employed and its great sensitivity to the shape
of the obstacle, i.e., the shape of the actual tire deformation. It may be
noted that the critical ratio h/a increases with increasing initial deflection
δ. This can be explained by the softening character of the nonlinear foun
dation characteristic. This nonlinearity will in general raise the critical
value of h/a.
TABLE 9.4.1. Critical values of h/a

h S a H
(cm.) (cm.) (cm.) «
2 1.7 9.8 0.205
3 2.1 11.2 .267
4 2.6 12.4 .322

Using this theoretical tire model, the variation of the vertical and longi
tudinal force has been calculated for a trapezoidal obstacle moving from
the leading edge to the center of the contact area. The results are shown in
figure 9.4.7. The influence of foundation nonlinearity and of shrinkage of
the tread-band are clearly demonstrated. The calculations have been car
ried out for a tire with a cross section proportional to that of the tire model
used in section 9.4.1, of which the dimensionless foundation characteristic
has been shown in figure 9.4.2. The tire model parameters are: r = 27 cm.,
δ = 1.5 cm., a ≃ 9 cm., b = 5 cm., l = 8.25 cm., hm = 8 cm.,pi = 2 atm. The
trapezoidal obstacle has the dimensions: height h = 2 cm., base line =12
cm., top line = 4 cm. This shape has been chosen for simplifying the calcu
lations during the first stage of contact. As the sloping side of the obstacle
touches the peripheral line at the instant of first contact, there will be prac
ANALYSIS OF TIRE PROPERTIES 737

tically no change in length of the circumscribed circle during the first stage
up to position 1. From this position on, reduction of the circumference will
take place. The dotted arc indicates the reduced circumference for the in
termediate position 2. The circumference is minimum as soon as the ob
stacle lies completely inside the circle.
For position 2 the pressure distribution along the contact line has been
shown. The vertical components qz are found with the aid of the stiffness
shown in figure 9.4.2. The integral of these components form the vertical
load - /•-. In the same way, the integral of the horizontal components qx
form the drag force -Fx.
In figure 9.4.7 the calculated force variations are shown. The vertical
force has been calculated for four different combinations of the following
assumptions: linear (cr =160 N/cm.2) and nonlinear foundation character
istic (fig. 9.4.2) and reduced and original circumferential length (curves fl,
b, c, and d). Curve b, computed using the linear theory but with length re
duction taken into account, is close to exhibiting a minimum. According
to condition (9.4.14) a minimum of curve b lower than the initial value
-Fzo would arise when the obstacle is reduced in size so that h < 0.076 ×
a = 0.68 cm. This critical value increases when the nonlinear characteristic
is taken into account.
The variation of longitudinal force does not show great sensitivity to the
model employed. Only curve e has been shown corresponding to the com
bination of length reduction and nonlinearity.
reduced circumscribed
c i re u msc r i bt d cirelt of tyrt
circlf • ( 2.3 ) (obstacle 0 or 1

softening found
ation characteristic

FIGURE 9.4.7. Calculated variation of vertical and longitudinalforce as afunction ofposition


of obstacle relative to contact center.
738 MECHANICS OF PNEUMATIC TIRES

The development of a linear response theory


One of the most important questions concerns the linearity of tire re
sponse to road irregularities. A linear behavior would simplify the analyti
cal treatment considerably. The tire has a finite contact length and con
sequently is subjected to a large number of inputs at the same time. The
contact length varies with obstacle height, and negative reaction forces
cannot be transmitted so that partial loss of contact can occur. In spite of
these difficulties, which do not benefit linearity, Lippmann and his associ
ates [16, 17] have experimentally shown that an almost linear relationship
exists between tire force variation and obstacle height. Tests have been
carried out with constant axle height and for the elementary step shape
obstacle. Even the measured force variation of a tire traversing a series of
composite cleats (combinations of blocks with rectangular cross sections
and extending over the full tread width) could reasonably well be derived
by linear combinations of the response to a unit step in elevation of the
road surface. A more refined combination of these responses, where a dis
tinction has been drawn between positive and negative steps, gave very
good correspondence to the measured total force variation. The appli
cability of this more refined method is limited since linearity has been lost;
for mathematical details we refer to the original papers [16, 17]. The linear
mathematical representation of the tire force response to road irregulari
ties, based on the principle of superposition enunciated in the fore-
mentioned papers, will be discussed hereafter. It has been stated that the
assumption of linearity leads to sufficient accuracy in most practical cases.
In figure 9.4.8 the positive directions are given for road coordinates (*,,
fc), wheel axle position or distance traveled (s = x), and reaction forces
(normal load -Fz and drag force -Fx). Figure 9.4.9 shows the force re
sponses to positive and negative unit steps in road elevation. The curves
correspond in character to data given in [17]. Note that the curves differ
slightly for positive and negative steps. The average of the absolute
changes in force may be taken as the characteristic response to a unit (up
ward) step. This function has been designated as i/\(£) for the longitudinal
force response and as ^z(0 for the vertical force response. The newly in
troduced coordinate £ denotes the distance traveled by the wheel axle from
the instant of first contact with the step irregularity.
We now decompose the road contour, Ic(xc), into an infinite series of
small positive and negative steps and use the principle of superposition in
order to obtain the total response. Expressed in the form of the super-

Xc Zc

FIGURE 9.4.8. Positive directions of coordinates and forces.


ANALYSIS OF TIRE PROPERTIES 739

FIGURE 9.4.9. Experimental force response to positive and negative steps in road elevation.

position integral of Duhamel, we obtain for the variation of the tire forces
as a function of the distance traveled by the foremost contact point s1 = s
+ a with a denoting the assumed! y constant half contact length:

(9.4.15)

At the starting position (s1 = a) the forces have the value F,zo. This general
integral approach of expressing the response may be written in the form of
the more widely applicable frequency response of tire forces to road irreg
ularities. Following Lippmann's analysis we first rewrite the above equa
tion. With the conditions
zc = 0 for xc<a
4v, = 0 for xc > s,
this equation may be written as follows

.4.16)

The contour of the road may be expressed as a Fourier integral over the
spatial frequency co^

(9.4.17)

The Fourier transform becomes now with the above version of the Fourier
integral:

(9.4.18)
740 MECHANICS OF PNEUMATIC TIRES

The quantity D(ωs)dωs/2ir is the contribution of those harmonic com


ponents of the road contour whose spatial frequencies are included in the
interval between ωs and ωs + dωs.
Differentiation of (9.4.17) yields

(9.4. 19)
S/

Substitution of this expression into (9.4.16) yields with £ = j, — xe:

17 (9-4.20)
L ~

The quantity in brackets is the Fourier transform of <!-„(£) describing the


spectral content of the tire force response to a unit step variation of the
road level. This function is designated as CXJ(ut). Hence eq. (9.4.20) be
comes:

(9.4.21)

Instead of measuring the response to a step, it may be more practical to


take a bump of short duration as the input. Such a bump may consist of
an upward step followed after a delay by a downward step of equal height.
The cross section of the unit bump has an area equal to unity and a4ength
tending to zero. The response to such a function simply equals the dis
tance derivative of &.,(£). The Fourier transform of the response to a unit
bump represents the frequency response function of the tire force to road
irregularities. This function, designated as HXJ(u,), consequently equals
iu,CfJ(u,).
We finally obtain for (9.4.21):

" #(«,)#-,(«>'"•'' «fc, (9-4.22)

In practice //„(«,) may be obtained by taking the Fourier transform of


the response to a bump which is short with respect to the shortest wave
length of interest and dividing the result by the area of the bump cross sec
tion.
Lippmann and Nanny [17] conducted a harmonic analysis of the re
sponse of vertical and longitudinal forces to a cleat by feeding the re
corded signals into a tuneable resonator (tire 8.15-15 two-ply, pi = 2 atm.,
deflection 2.54 cm.; cleat 1.27 cm. long, 0.63 cm. high; speed ≃ 1 m/sec.).
The amplitude spectra exhibit two broad maxima. For the vertical force
response their maxima occur at the 2.4th and 14th order of the tire revolu
tion, and for the longitudinal response at the 8th and 21st order. These fig
ures correspond for the tire under consideration to the following frequen
cies per unit of speed, reciprocal value of wavelength expressed as (cycles/
sec/meters/sec.): vertical 1.1 and 6.3 c/m., longitudinal 3.6 and 9.4 c/m.
At shorter wavelengths (higher frequencies) the amplitude decays to low
values.
ANALYSIS OF TIRE PROPERTIES 741

In the theories and experiments mentioned above the motions are as


sumed to develop slowly so that inertia forces can be neglected. The trav
ersing of obstacles at higher velocities involves high frequency vibrations
and dynamic forces in the tire tread band, the nature of which will be
dealt with in section 9.4.2.
In the literature, tests are described which have been carried out at
higher speeds. In these tests the wheel axle has been free to move in its
suspension. Care must be taken in the interpretation of the results as they
do not solely refer to the response of the tire. We may mention here the
work of B arson [18] and Guslitzer [19] who measured the motion of ve
hicle components when passing over an obstacle. In the latter reference at
tention has also been given to failure of tires due to obstacles. The forces
acting from the tire to the ground when rolling over an obstacle have been
measured by Hey [20] for a relatively long single sinusoidal obstacle and
by Senger [9] for a relatively short semkylindrical obstacle. The velocities
in these latter two tests did not exceed 50 km/hr.

Longitudinal Slip
A tire which rolls freely at constant speed of travel, i.e., not subjected to
driving or braking torques, requires a thrust in the longitudinal direction
acting on the wheel axle in order to overcome the rolling resistance. As has
been shown in figure 8.2.56 the rolling resistance depends on the normal
load. For steady state motion the thrust P is in equilibrium with the drag
force Fr and the rolling resistance moment Mr = rf, where r1 denotes the
axle height above the road surface. At constant speed we may employ the
linear relationship
MA=rfr) = D,W= CS (9.4.23)
with W(=— Fz) denoting the normal load and δ the normal tire deflection.
The coefficients of rolling resistance Dr and Cr are, at least in the low speed
range, not influenced much by the speed V (cf. figs. 8.2.57, 8.2.60).
At free rolling the angular velocity fl and the speed V are related
through the effective radius of rolling r,,
V = Or.. (9.4.24)
It has been found experimentally that
r, < r, < r. (9.4.25)
According to Whitbread (cf. [1, 13 or 21]) the following approximate rela
tion exists
r, = r - ViS. (9.4.26)
This expression has been found by considering the compression of the
tread band in the contact area due to the normal deflection of the tire.
Tests with aircraft tires roughly confirm this formula [13]. For a tire with
an inextensible thin tread band and a worn-off tread pattern, so that creep
of the tread band with respect to the road does not take place, the effective
radius tends to the free radius r. Figure 8.2.43 provides more information
about the variation of rolling radius with speed for different tire construc
tions.
We define now the longitudinal or tangential creep or slip velocity Vcx
as the tangential speed of a point C fixed to the wheel rim and situated at
road level (fig. 9.4. 10a). At the instant considered point C coincides with
742 MECHANICS OF PNEUMATIC TIRES

the contact center C, which, according to its definition, is located on a line


normal to the road surface and passing through the wheel axle. Con
sequently, for a horizontal road surface, C lies directly below the wheel
axle.
The longitudinal slip value relates the slip velocity to the speed of
travel.*

«c--F«/F- — (9.4.27)

In section 9.5.1 a slightly different definition for the longitudinal slip value
has been adopted, viz κ = — Vcx/rlΩ. Their values are nearly the same at
moderate longitudinal slip.
Assuming that relation (9.4.26) exists, the longitudinal slip speed be
comes at free rolling
(9.4.28a)
or, expressed in more general terms,
Vcf = 7j'«2 (9.4.28b)
from which we obtain the longitudinal slip value at free rolling

(9.4.29)

• In some texts, instead of* defined in eq. (9.4.27) the symbol - j is used. Furthermore, rl may be replaced by rr The slip
speed Vc x would then be defined u the longitudinal velocity of point P (attached to wheel rim) shown in fig. 9.4.10 (cf
[44]) In that case, at free rolling s = 0 and at wheel locking s = 1.

of rotation
FIGURE 9.4. 10a. Rolling resistance force Fr and moment Mn longitudinal slip velocity Va
and center of rotation P at free rolling.
ANALYSIS OF TIRE PROPERTIES 743

radial steel 7mm


cross ply 7
cross ply 0
redial steel 1

3600 N

10 15 20 26 30
(mm)

FIGURE 9.4. lOb. Measured decrease of effective rolling radius (difference offree radius r and
effective rolling radius rj as a function of tire deflection 8. The mathematical representation
around the nominal wheel load of 3500 N, according to eg. (9.4. 26a), has been indicated.
Two different types of tires are used at full tread depth (7 mm) and at almost or completely
bald conditions. The speed of travel V amounts to approximately 160 km/h. Its influence
(20-160 km/h) appeared to be small. At low deflections S, the center ofrotation P appears to
be located above road level, especially for radial tires.

The value of TJ' is approximately % for bias ply tires and tends to unity in
case of an inextensible tread band with little tread rubber.
The location of the center of rotation of free rolling, point P in figure
9.4. 10a, is located at a distance tj'S below road level. Measurements in
dicate that in fact TJ' is not a constant but varies with deflection δ and may
even become negative below small values of δ. Instead of defining the lo
cation of the center of rotation of the rim with respect to the road level we
may choose an imaginary effective level a distance Δ above the real con
tact surface. This effective rise Δ may account for the influence of tread
pattern and depth and may range from almost nil (bare tires or continuous
744 MECHANICS OF PNEUMATIC TIRES

ribs) to approximately the depth of the tread grooves (isolated studs). The
height of the center of rotation below the effective road level may be rea
sonably described with ijS. The fraction 17 is considered as a constant.
Hence, the following expression for TJ' is obtained:
Tj' = Tj - A/8 (9.4.30)
so that eq. (9.4.26) may be written as follows
r. = r - (1 - Tj)8 - A. (9.4.26a)
Since it is rather obvious that re tends to r when δ nears zero, it is expected
that at very small δ the effective rise Δ can no longer be considered as con
stant but will vanish with δ. This has been experimentally verified (cf. fig.
9.4. 10a). In the range of deflections considered, Δ may be assumed con
stant. The tire parameters TJ and Δ can be estimated from experimental
data. Measurements on a 3m diameter drum gave the following typical re
sults. Radial steel belted 14" tire, 7 mm tread depth: TJ = 0.90, Δ = 10.5
mm; 1 mm depth: TJ = 0.97, Δ = 6.0 mm. Nominal load W = 3500 N.
Cross-ply 14" tire with ribs, 7 mm tread depth: TJ = 0.61, Δ = 3.6 mm; bald
tire: TJ = 0.54, Δ = 1.6 mm. Nominal load W = 3500 N. After introducing
the new parameters TJ and Δ, equation (9.4.29) becomes:
K ~ _,,'*=_,, i + A . (9.4.29a)
r r r
When applying additional driving or braking torques the slip value
changes in a manner analogous to that of the tire cornering characteristic
relating lateral force to slip angle. Figure 9.4. 1 1 gives an example of mea
sured longitudinal slip characteristics, taken from Hörz [22].
Longitudinal slip is based on the following mechanisms. The first stems
from the inextensible band theory, which states that since the same quan
tity of material passes at the top and at the bottom of the tire per unit time,
longitudinal slip must occur at the road level. Secondly, we have the influ
ence of compressibility or extensibility of the tread band. Compression in
the contact area due to vertical deflection of the tire diminishes the slip
due to the first mechanism. Also under the action of tangential force, the
density in the neighborhood of the contact area may change. As a result of
this, the tangential velocity at the top and the bottom do not necessarily
have to be the same. Third, one may include the influence of tread rubber
longitudinal elasticity, which enables the tread band to move with respect
to the road without sliding. Finally, we have the possibility of partial or
total sliding in the contact area.
The first effect merely depends on geometry and does not involve result
ant longitudinal forces. The second and third effects give rise to pure de
formation slip characterized by the longitudinal or tangential creep, or slip
stiffness C« defined by the relation

(9.4.31)

Since the tensile rigidity of the tread band is relatively large, the deforma
tion slip is expected to be mainly due to tread rubber elasticity. The longi
tudinal slip stiffness Cκ can then be expressed in terms of longitudinal
ANALYSIS OF TIRE PROPERTIES 745

<o 60 90 '/. tOO

brake slip value -K


FIGURE 9.4.11. Example of measured longitudinal slip characteristics [22].

tread profile rubber stiffness per unit area cpx, half contact length a and
half contact width b of the assumed rectangular contact area:
(9.4.32)
This expression suggests that Cκ varies approximately as the square of the
contact length. With the assumption that the contact length varies quad-
ratically with the normal deflection δ, we obtain for a linear tire load-de
flection characteristic W = Czδ:

(9.4.33)

with index o denoting the original situation. From test data [13, 21, 22, 23]
it can be deduced that the value of Cκ ranges from 5 to 15 times the verti
cal load W. This indicates that Cκ is of the same order of magnitude as the
cornering force stiffness CFα (sec. 8.5.1).
For small deviations from the steady state conditions the longitudinal
force-slip characteristic may be linearized and written in the form
F, .+£. (9.4.34)
where the index o indicates the original situation and the upper bar in
dicates the deviation from this situation. The variable part /\ is a linear
function of the variation of vertical deflection 8 and of longitudinal slip
value k. With -Fxo = Fro denoting the original rolling resistance force we
obtain in general:
Fx = -^ + ^S + ^ic. (9.4.35)
746 MECHANICS OF PNEUMATIC TIRES

The values of the partial derivatives should be determined experimentally.


From eq (9.4.31) the value of dFJdK obviously equals Cκ0. With our
knowledge of effective radii at free rolling we may derive BFJdd theoreti
cally. From figure 9.4.12 we deduce for the longitudinal force at W =
W, + »P(or S = δ0 + 5) and κ = κ0 + κ:
FM--F, (9.4.36)
The quantity κ′ denotes the variation in longitudinal slip value of a freely
rolling tire due to a variation in normal deflection 8. According to eq
(9.4.29a) the value of κ′ reads:

K' - ij — • (9.4.37)

The rolling resistance force Fr becomes according to eq (23):

(9.4.38)

We finally obtain for eq (9.4.36), linear in the variations a and 8, and


with δ neglected with respect to r:

- Fro (9.4.39)

Herewith, the coefficients of eq (9.4.35) have been determined. The equa


tion above is of great value for the theoretical assessment of the longitudi
nal force response to road irregularities and tire nonuniformity. The next
section is devoted to this problem.
At relatively large values of longitudinal slip linearity is lost. For the

.W
Slope CK
Free rolling
oriqinol condition
ot load M,

-Free rolling at
load W

FlOURE 9.4. 1 2. Longitudinal slip characteristics near the steady state conditions (subscript o).
ANALYSIS OF TIRE PROPERTIES 747

theoretical treatment of the behavior in the nonlinear range it is more


straight forward to use the alternative definition of longitudinal slip value
κ = — VCJ$LT, which results in a complete analogy with the side slip phe
nomenon. For more information we refer to side slip theories enunciated
in section 9.5.1. The latter part of this section treats the problem of the in
teraction of longitudinal and lateral slip. We furthermore refer to the early
theories of Reynolds [24], Fromm [25], and Julien [26]. Experimental data
are available in the publications of Yu Chen [21], Hcirz [22], and a very
extensive investigation of the behavior on wet roads by Holmes and Stone
[23]. For aircraft tires we refer to Smiley and Home [13].
Theory of the Longitudinal Force Response to Road Wavyness, Tire Non-
uniformity and Vertical and Longitudinal Axle Motions
For the development of a theory of the longitudinal force response to
in-plane variations of wheel axle position, road contour, tire radius and
tire radial stiffness, a theoretical tire model is proposed, shown in figure
9.4.13, in which the following properties are represented:
1. radial tire elasticity with stiffness normal to road Cz,
2. tangential carcass elasticity with stiffness tangential to the road C»,
3. tangential slip of carcass with respect to the road with slip stiffness
Cκ,
4. rolling resistance moment Mr with coefficient Cn
5. moment of inertia of rotating mass including a great portion of the
tire, I,.
The system has two degrees of freedom, viz. the angle χ denoting the
deviation of the angle of revolution from the steady state situation charac
terized by the constant angular speed SI, and the torsion angle if of the
wheel relative to the lower tread region (see fig. 9.4.13). The input vari
ables are:
1 . horizontal and vertical wheel axle motions xa and za,
2. radial runout r and radial stiffness variations Cz,
3. road level variations z,.
It may be noted that in reality the tire nonuniformity input quantities may
vary with rotational speed Ω. This effect will not explicitly be taken into
account. The inputs are assumed to be of relatively long wavelength so
that first, the curvature of the road contour is much less than that of the
tire; second, the wavelength is much longer than the contact length be
tween tire and road, which reduces the treatment of the longitudinal slip
behavior to the quasistationary problem; third, the frequency is small rela
tive to the first natural frequency of the tire body (40-50 Hz) permitting
tire inertia effects to be neglected.
We shall restrict ourselves to the investigation of the response to the in
put variables mentioned above. It may be noted, however, that other input
quantities may influence the longitudinal force. Among these, the nonuni-
formly distributed mass along the tire circumference is also important.
One effect of nonuniformity in mass distribution may be the variation in
effective normal stiffness with speed of travel. By measuring the periodic
tire radial stiffness variations as a function of rotational speed this effect
will be included automatically. Also, the contribution of the mass nonuni
formity to the change of radial runout with rotational speed should be rec
ognized.
The mass of the lower portion of the tire is subjected to the action of
tangential Coriolis forces which arise when the lower tread-band portion
748 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4.13. Tire model for assessment of longitudinal force response.

moves in a radial direction with respect to the wheel axle. Similarly, a var
iation in tread-band mass causes a resultant tangential Coriolis force even
in the case of fixed axle height. In both cases the variation in angular mo
mentum of the tire-wheel combination about the wheel axle gives rise to
the fore and aft fofec variations.
The influence of these effects deserves closer investigation. It may be
noted, however, that the whole phenomenon is restricted to a portion of
the tire near the contact area. The tangential forces are in equilibrium
with the inertial forces acting on the particles present in this portion.
Therefore, the net force to be measured at the wheel axle is expected to be
very small, if present at all.
A rough calculation may provide some quantitative insight. Consider a
tire with radius r = 0.3 m., contact length 2a = 0.15 m. and tread-band
mass per unit length 1 .5 kg/m. The mass which moves when deflecting the
tire becomes approximately 0.25 kg. When the tire rolls over a flat surface
at a constant speed V = 30 m/sec. and a constant angular speed S2 = 100
rad/sec., and moves vertically with a frequency of 10 Hz and an ampli
tude of 0.005 m., the lower tread band portion will be subjected to a Co
riolis force which has an amplitude equal to
2 X 0.25 X 2ir X 10 X 0.005 X 100 = 16 Newtons = 3.57 Ibs.
The problem of the influence of a gradient in mass distribution may be il
lustrated with the following example. Consider the same tire in steady
state motion but possessing a heavy spot of 0.05 kg. This tire element will
be subjected to a Coriolis acceleration at the entrance of the contact area
equal to 2a(V/r)2. The Coriolis force consequently becomes 0.15 × 1002 ×
0.05 = 75 Newtons = 16.8 Ibs. Theoretically, these forces are measureable
at the contact surface and not at the wheel axle.
ANALYSIS OF TIRE PROPERTIES 749

We shall now turn to the restricted problem where inertia effects in the
lower region of the tire are not considered. For the tire model of figure
9.4.13 the equations of motion will be presented. Again, the symbols pro
vided with index o denote the stationary or average value and the upper
bar indicates the variable part. We restrict ourselves to small deviations
from the rectilinear stationary motion, and can therefore linearize the
mathematical representation.
The vehicle has a constant speed of travel V. The distance traveled be
comes
s = Vt. (9.4.40)
The contact center has a coordinate in x-direction somewhat deviating
from s.
We write
xc = s + Ax (9.4.41)
where

Ax = x0-r,-^- (9.4.42)
mXe

Taylor's formula yields for the slope at xc

-L.^*^... (9.4.43)
dxc ds ds1
In the linear representation we may omit the terms with Ax. For the longi
tudinal force acting in the contact center we obtain
X.-- Frm (9.4.44)

X=F,+ W0^- (9.4.45)

The tangential force variation becomes with eq (9.4.39) and with the con
tribution of a change in Fr due to a variation in Cr — (Cr> + C,) resulting
from load variations directly caused by radial stiffness variations C,\

F, = cjc + i {(T, C.. - C,)S - «.C) . (9.4.46)


*o

This equation, valid for steady state longitudinal slip, can be used with
good approximation for nonstationary motions with wavelengths which
are long relative to the contact length.
The tangential slip value denned by eq (9.4.27) becomes

(9A47)
with the steady state value

Ko - — =- s -if (9.4.48)
750 MECHANICS OF PNEUMATIC TIRES

and variable part

(9A49)

The tangential slip speed reads


V« - V + xa - (Q + x + for,. (9.4.50)
With
r, = r - 8 - rto + f - 8
we obtain for the variation of K«
?„ - *. - Q01 ~ *) - rto(x + <£)• (9.4.51)
When neglecting 6 with respect to r we obtain for eq (9.4.49)

£=- {*„ - B(r - 8) - r0(x + *)} . (9.5.52)

The equation of constraint for 8 reads:


8 - - ze + za + r. (9.4.53)
For the variation of the tangential force we have the additional relation
F, - - C«r.*. (9.4.54)
The normal load
W = C,8 = (C,0 + C,)(80 + 5). (9.4.55)
Instead of using the variable stiffness r. we prefer to introduce the vari
able static deflection S,,. This quantity can, in principle, be determined by
measuring the variation of the axle height of a perfectly round tire which
rolls under constant vertical load Wa over a smooth flat surface. The fol
lowing relation applies
(9.4.56)
Hence, the following relation holds for the variation in normal load
^-C^8-8.). (9.4.57)
According to relation (9.4.23) the variation in rolling resistance moment
becomes
CJ + 80C, - 8.) (9.4.58)
with Cn and C, given by

C ro (9.4.59)

In eq (9.4.58) the relatively small influence of the longitudinal shift of the


point of application of the vertical load W due to out of roundness f has
been neglected.
ANALYSIS OF TIRE PROPERTIES 751
The dynamic equilibrium about the wheel axle is governed by the dif
ferential equation
'*X = - FS, ~M,= Fn(f - S) - - M,. (9.4.60)
For the eight unknown quantities χ, Fx, Mn \, \l/, o, ic and s we have at our
disposal the eight equations (9.4.40, -45, -46, -52, -53, -54, -58, -60). For
the unknowns X, \ and \j/ we obtain for the elimination of the remaining
unknown quantities the following set of equations:

(9.4.61)

Cro)(zc - z.) - (9.4.62)

*•-«- — f + x.

C.
(9.4.63)
c.
These equations of motion can be used in the analysis of wheel suspension
vibrations. We shall concern ourselves here with the derivation of the re
sponse functions of the longitudinal force to each of the input quantities.
It should be recognized that some of these input quantities depend upon
each other in a way which is a function of the suspension properties.
A final elimination of the variables χ and ψ and the introduction of the
operator p = d/dt yields the following expression for the response of the
longitudinal force variations A' to the road irregularities z,, the radial run
out f, the variation in static deflection $„ and furthermore to the longitudi
nal and vertical axle motions x,, and z,;

X-- {/Jl - p + ra(Fro CJ\ (z. - fe)


I I J

C
A" + -pzc (9.4.64)

in which

From this expression of combined responses the single absolute frequency


response functions can be derived.
Frequency response to vertical axlr motions
After replacement of p by /«, with ω denoting the frequency of the mo
tion, we find for the response to vertical axle motions the following ampli
752 MECHANICS OF PNEUMATIC TIRES

tude ratio,

(9.4.65)

where have been introduced the nondimensional quantities

(9A66)

(9467)

(9.4.68)
W*

with w denoting the rotational natural frequency of the standing tire


about the wheel axle

(9.4.69)

For illustration, some aspects of the response will be discussed for a tire-
wheel combination with parameter values:
Ix = 0.6 kgm.2, Cef = 5 X 103 N*/m., ra = 0.3 m.,
nxo = uxo/2ir = 43 Hz, TJ = %, C.0 = 40000 N, (9.4.70)
F,a = 60 N, 6,, = 0.02 m., W0 = 4000 N, Cn = 900 N.
The dimensionless parameters become, for this configuration,
/, = 0.023 V and /2 - 0.342 K
For w = 0 we find
X
3200 N/m. (9.4.71)

For small w («: wxo) we may simplify to


X

3200 1 1 + ( 1 + 0.233 V2) -j^-A , (9.4.72)

• N - Newton - .223 lb«. fora.


ANALYSIS OF TIRE PROPERTIES 753

where V denotes the speed of travel expressed in m/sec. and n denotes the
frequency of the axle motion in Hz. It may be observed that the magnifi
cation of the zero frequency response (9.4.71) is in particular due to the
large tangential slip stiffness C.0 and the moment of inertia Ix. The effect
of changing Ccx is noticeable only at very small values of speed V. The fac
tor T) appearing in eq (9.4.29a) can exert considerable influence. A value of
7; close to one is favorable, which can be explained by the fact that if tj = 1
the variation of the effective radius at free rolling with deflection becomes
very low and is in that case only due to the change in rolling resistance.
At higher frequencies the amplitude ratio deviates from the approxima
tion (9.4.72). In figure 9.4. 14a response curves are drawn for several speeds
of travel. Besides the point at zero frequency where the curves come to
gether, two more invariant points appear to exist where the response is in
dependent of V. Since these points show the same amplitude ratio and are
located near the natural frequency nxo, their height gives a good indication
of the maximum amplitude ratio, which occurs near the natural frequency
and is virtually independent of V. The curves show furthermore that for
the higher range of speeds this maximum becomes very broad, which
means that high amplitude ratios can- be attained at relatively low fre
quencies.
From eq (9.4.65) we obtain for the amplitude ratio in the invariant
points

or a magnification of about 15 times the zero frequency response. From


the expression above it follows again that large CM and low 17 are unfavor
able. The speed V and the parameters Ix and Cex do not influence the re
sponse in the higher frequency range, near the frequencies of the invariant
points, given by

(9.4.74)
/,
Obviously, the invariant points lie close to and on either side of the natu
ral frequency.
With both cross-ply and radial-belted tires, tests have been conducted to
assess the frequency-response curves experimentally. A special test-stand
has been used with which forces acting on the vertically vibrated wheel
axle can be measured up to an excitation frequency of about 24 Hz (cf. fig.
9.4.14b). Forces are measured in three perpendicular directions at both
ends of the axle by a set of piezo-electric elements. The structure is moved
up and down by a hydraulic actuator. The tire rolls over a drum of 3 m
diameter. The measured and calculated response plots are shown in figs.
9.4. 14c and d (cf. [47]), using eq. (9.4.65). The diagrams demonstrate that
the theory provides a satisfactory description of the real tire behavior.
Mainly due to the lower value of rj for the cross-ply tire, the amplitude re
sponse of this tire turns out appreciably higher.
754 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4. 14a. Frequency response curves for longitudinal force X with respect to vertical
axle motions i,, calculatedfor the tire-wheel parameter values (9.4.70).

Frequency response to longitudinal axle motions.


From eq (9.4.64) we find for the response to xα:
X C..F1
(9.4.75)

This is the well-known form for the acceleration response of a single mass-
spring system excited by a force. The dimensionless coefficient of damping
becomes for our system:

(9.4.76)

FIGURE 9.4.14b. Side andfront view of test-rig for the assessment of wheelforce response to
vertical wheel axle oscillations.
I . hydraulic actuator, 2. piezo-electric force transducer. 3. elastic hinge
ANALYSIS OF TIRE PROPERTIES 755

CROSS-PLY EXPERIMENT RADIAL- PLY


STEEL-BELTED
P 1 a f:.l
40
LONGITUDINAL
FORCE PER UNIT
VERTICAL AXLE 30
MOTION

20 11.8m.-s

10
90'
PHASE LAG
OF LONGITUDINAL
FORCE WITH
RESPECT TO ,j
VERTICAL AXLE '20
MOTION

FIGURE 9.4. 14c. Measuredfrequency response curves (amplitude and phase lag) oflongitudi
nal force X (or F, in this case) with respect to vertical axle motions zr Wheel load W0 -
3000 N; tread depth 6.5 mm.

Damping obviously increases with increasing speed V and decreases with


increasing slip stiffness Cκo. Compared with the response to z., the speed V
has an opposite effect whereas a variation of C.,, changes the response in
the same direction. The effects of Cc x and / are more difficult to evaluate
since they influence the response also in other respects. At relatively low
frequencies Ix is the most important parameter. For the parameter values
of (9.4.70) we find a dimensionless damping coefficient f1 = 0.023 V. Con-

(CROSS PLY) THEORY (RADIAL-PLY


mn^ V= 27.6^ STEEL-BELTED)

40

LONGITUDINAL
FORCE PER UNIT
VERTICAL AXLE
MOTION
20 11 Bm;s

20 Hi 20 Hz
90'

PHASE LAG
OF LONGITUDINAL
FORCE WITH 120'
RESPECT TO
VERTICAL AXLE
MOTION
150'

FIGURE 9.4. 14d. Calculatedfrequency response curves (amplitude and phase lag) of X with
respect to z. Parametersfor cross-ply tire/radial belted tire: Fn - 60/25 N; r, - 0.3/0.294
m; Cn - 900/335 N; Ccf - Kf/565000 N/m; /_ - 0.8/0.8 kgm2; C,0 - 45000/80000 N;
H - 0.6/0.85; nxo - 53/39.5 Hz.
756 MECHANICS OF PNEUMATIC TIRES

sequently, the damping becomes critical at a speed approximately equal to


V = 40 m/sec.
Frequency response to variations in static deflection
From the expression for the combined response (9.4.64) we obtain
1+4/.V
-J2 (9A77)

which corresponds to the displacement response of a single mass-spring


system excited by support oscillations. The response to variations in radial
stiffness or static deflection as formulated above is very small. However,
the indirect response with axle free to move vertically will be appreciable.
The variations in radial stiffness induce vertical axle motions which in
volve a response of the longitudinal forces just discussed.
Frequency response to radial runout
The absolute frequency response function to radial runout f reads:

- !'
^

with the dimensionless parameters

(9.4.79)

(9.4.80)

It may be noted that great similarity exists with the response to vertical
axle motions (9.4.65). The level of the response, however, can become con
siderably higher in the higher speed range. This is due to the fact that in
the dimensionless parameter f3 the factor 1 — TJ has been replaced by —TJ.
For tires with ?/ tending to one (inextensible tread band and little tread
rubber) which was favorable for the response to vertical axle motions, it
turns out that their response to radial runout becomes worse. For the val
ues used in (9.4.70) we obtain for the amplitude ratio at the invariant
points

85900 N/m. (9.4.8 1)

or about 8 kgf. per mm. radial runout. The magnification with respect to
the zero frequency response amounts to approximately 27 times.
Experiments with eccentrically mounted wheels showed frequency re
sponse characteristics which were in excellent correspondence with theo
retical predictions (cf. [47]). In these tests, of course, the speed V changes
proportionally with the frequency w.
ANALYSIS OF TIRE PROPERTIES 757

Frequency response to road irregularities


According to eq (9.4.64) the response to road irregularities can be
formed by adding an extra term to the negative of the response to vertical
axle motions

The solution for this combination of responses to variations in f, and in


slope dzjds is difficult, and is in fact impractical since no useful informa
tion can be obtained other than that which can be drawn immediately
from (9.4.82). One might as well consider the slope dzjds as an isolated
input, the response to which must be regarded in combination with the re
sponse to axle height to roadway variations. The response to slope varia
tions at constant axle distance from the road surface simply equals the av
erage normal load W,,.
9.4.2 High Frequency Properties
In the high frequency range, well above 30 Hz, tires exhibit a number of
natural frequencies in vertical (radial), longitudinal (tangential) and lat
eral direction and will consequently snow continuously distributed vibra
tions of side walls and tread band.
Although the intensities of the high frequency vibrations of the un
sprung masses are lower than those of the lower frequency resonances, re
search on this class of high frequency vibrations is valuable since they are
lightly damped and in some cases are even amplified because of local reso
nances occurring in the vehicle. They also play a fundamental role in
acoustical effects.
The high frequency vibration behavior of rolling tires is a very com
plicated matter and has been dealt with only briefly in the literature. In
the following, important experimental and theoretical investigations on
high frequency tire properties will be discussed.
Vertical and Longitudinal Vibration Transmission
In recent years the transmission of high frequency vibrations from con
tact patch to wheel hub has been experimentally investigated by Chiesa,
Oberto and Tamburini [27, 28] and by the team of Gough and Barson [18,
29].
Chiesa and his group mainly reported on investigations with the non-
rolling tire excited by a vertically vibrating platform on which the tire
stands. Barson and Gough, in addition, executed experiments with tires
which roll over a drum to which can be attached a great variety of arti
ficial road irregularities, ranging from sinusoidal shapes to random sur
faces. In the case of a sinusoidal profile a wide frequency range can be in
vestigated by varying the drum speed. In their extensive investigation the
longitudinal aspect of vibration transmission has been included.
Figures 9.4. 15a, b (from [18]) respectively show the vertical transmission
ratio (amplitude ratio) of hub relative to platform for the nonrolling tire
and the r.m.s. value of the vertical hub acceleration for the tire rolling over
a sinusoidal drum surface with wavelength 0.133 m. and amplitude = 3
758 MECHANICS OF PNEUMATIC TIRES

mm. Both tests have been carried out with a laboratory strut-type wheel
suspension showing vertical and longitudinal compliance, the latter result
ing in a longitudinal natural frequency of about 11 Hz. In figures 9.4. 16a,
b the results of analogous tests for longitudinal vibration transmission
have been shown. In this case the vibrator platform is excited in the longi
tudinal direction, while for the drum test the same sinusoidal surface is
used as in figure 9.4. 156.

FIGURE 9.4.15. Comparison of cross-ply and radial-ply tires in their vertical vibrationai
behavior. (For values along the abscissa see Fig. 9.4.16).
Test carried out on vertically vibrating platform (a) and on rotating drum (b) with sinusoidal surface (wavelength 0.133
m, amplitude 3 mm. speed 5-100 km/h) [18].
ANALYSIS OF TIRE PROPERTIES 759

«o eo BO 100 200 3OO


FREOUESO — HI

10
LONGITUDINAL DRUM

' fv
CROSS- «.Y TYUt

30

10 00 OOA01OO JOO 300


FRCOUCNCr-HI

FIGURE 9.4.16. Comparison of cross-ply and radial-ply tires in their longitudinal vibrational
behavior.
Tesls carried out on longitudinally vibrating platform (a) and on rotatini drum (6) under lame condition.* of flfure

These experiments have been carried out with radial and bias ply tires
of the same dimensions. For clear comparison of the properties of these
tires their response ratio has been shown in the lower diagrams. It may be
noted that for the vertical transmission of vibrations qualitative similarity
exists between results of both types of test, platform and drum. The same
resonance frequencies (attributable to both tire and suspension) arise, and
from both tests the advantage of radial ply tires appears above 100 Hz, the
760 MECHANICS OF PNEUMATIC TIRES

"road roar" range. Their disadvantage in the range of 60-100 Hz is also


clear. In particular, for the longitudinal aspect of vibration transmission,
the drum test appears to be essential. From this test (fig. 9.4.166) it can be
shown that the radial ply tire causes a much greater longitudinal response
below 100 Hz than the cross ply tire.
The similarity of vertical response between rolling and standing tires led
Chiesa to adopt the oscillating platform test for further detailed investiga
tion of the vertical vibration transmission. In order to overcome the prob
lem of the presence of resonances other than those attributable to the tire,
the tire has been mounted on a massive heavy wheel of approximately the
same weight as the load sustained by a single tire fitted to a car. This
wheel was supported on the vibratory table by an air spring system of neg
ligible stiffness.
In the experimental investigation described in [27] a comparison has
been made between the vibrational behavior of radial and cross ply tires.
The oscillating platform produces vertical hub oscillations via tread band
and sidewall vibrations distributed along the tire circumference. The ques
tion of whether a correlation exists between these distributed vibrations
and the complex transfer function of platform motion to hub motion
forms the major part of that research.
The amplitude ratio of the hub oscillation and the sinusoidal platform
oscillation has been shown in figure 9.4.17 for both types of tire as a func
tion of platform frequency. With the radial tire the system exhibits a very
pronounced resonance at about 90 Hz, followed by a number of less im
portant resonances. The cross ply tire, however, exhibits only one reso
nance near 150 Hz. In practice, this indicates that the radial tire has more
potential for developing resonances in the vehicle than the cross ply tire.
The tests have been carried out with an inflation pressure p,= 1.5 bar, (1
bar ≃ 1 atm.). At higher inflation pressures the resonances appear to shift
to slightly higher frequencies and the amplitude ratios increase a little.
This is true for all resonances except the peak occurring at about 230 Hz,
being due to the resonance of the air column within the tube. This appears
to vary only with tire size and type of gas used for inflating the tire. Expe
rience shows that the whole picture of the transmission curve changes sub
stantially when varying the tire size. When the type of tire construction re
mains the same, the only effect is that with larger sizes all resonance
frequencies shift towards lower values. These results are of importance for
the tuning between tire and vehicle.
For determining the correlation between transmission curves and tire
vibrations, Chiesa and his co-workers found the shape of the deformed
equatorial line by measuring the hub and tire tread band and sidewall mo
tions by means of accelerometers. The instantaneous displacement of a
point of the tread band is obtained by vectorial combination of the radial
and tangential displacements deduced from the signals of micro-
accelerometers placed between tube and tire. By connecting the positions
of the points at several locations (azimuth angles) the instantaneous defor
mation line is obtained. These displacements occur relative to the station
ary equatorial line, which is deformed due to the average vertical load.
For purposes of this research it has been considered acceptable to take as a
reference equatorial line a perfect circle. In figure 9.4.18 deformation pat
terns of a radial and a cross ply tire have been shown with respect to the
circular reference line. The deformation patterns which are shown here for
a series of excitation frequencies are obtained at that instant where the
ANALYSIS OF TIRE PROPERTIES 761

50 100 150 200 Hz

FIGURE 9.4.17. Amplitude ratio between hub of heavy wheel and vibrating platform [27].

highest vertical resultant at the hub of the heavy wheel occurs. The phase
delay angle <f>, of the hub vertical motion relative to the shake table motion
has been indicated for each frequency. The vertical bold line segment in
dicates the amplitude of the platform motion.
At low frequencies the deformation is limited to the region near the
contact area. By increasing the frequency the first mode of vibration is at
tained. At this natural frequency (radial ply ~ 90 Hz, bias ply ~ 150 Hz)
the deformation patterns are similar (two nodes). It can be seen from the
figure that with increasing frequency the number of wavelengths along the
circumference increases. At a frequency of 190 Hz the radial ply tire al
ready has seven wavelengths whereas the conventional bias ply tire shows
only two wavelengths at this frequency.
From observations over one complete period of excitation it has been
found that the wavy deformation as a whole does not vanish periodically.
The nodes appear to move a little along the circumference, which is pecul
iar to dissipative distributed vibrations. Moreover it appeared that the
number of nodes may vary during one period.
Measurements of the lateral displacements of the sidewall indicate that
a simple, generally valid relation with the radial tread displacements ex
ists. It appears that these motions are always out of phase, i.e., when the
one contracts the other expands. This is true except in the direct vicinity of
the contact area where both motions are in phase.
For investigating the correlation between the deformation line of the
tread band and the vertical force transmission, the following procedure
was adopted. In increments of 10 degrees of azimuth angle the vertical
component of the displacement of the points of the equatorial line have
been determined, after which their algebraic sum has been calculated.
This sum has been determined for a number of instances over one ex
citation period and it appears that the variation with time is approximately
762 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4.18. Deformation lines of radial and cross-ply tires for a series of frequencies
taken at the instant ofhighest vertical resultant characterized by the phase delay angle $, rel
ative to the platform motion.

harmonic. In figure 9.4.19 the amplitude ratio and the phase relationship
of this calculated sum with respect to the platform motion has been
shown, together with the measured response of the hub motion. It appears
that the agreement is very good between the calculated sum of the vertical
deformation, which may be considered to be proportional to the force
transmitted, and the experimental results of figure 9.4.17, after the latter
are cleared of low-frequency components. As Chiesa states, the correlation
which has been found to exist furnishes a broad basis for the understand
ing of the high-frequency (vertical) behavior of tires, to add to the well
known low-frequency phenomena.
Similar research can in principle be carried out for the investigation of
longitudinal and lateral force transmission. It may provide some insight,
but the practical application of the results obtained with a horizontally
shaking platform will be difficult, since the motions of the lower part of
the tread-band running on a road are not known due to longitudinal and
lateral creep phenomena. Moreover, for longitudinal force transmission
the slope of a wavy road surface plays an important role, as has been
shown in the previous section 9.4.1.
The natural frequencies and the mode shapes of the motions in these
two directions however, are no doubt of importance for general informa
tion and for purpose of tuning the tire-vehicle combination. The informa
tion shown in figure 9.4.20 has been taken from reference [18]. It shows
several modes of the tire-suspension combination. The tire is of the radial
ply type and the suspension is of the strut type having large longitudinal
compliance. The resonances will change with mass, polar moment of iner
tia and longitudinal compliance of the tire. Probably the only peak which
is due to tire natural vibrations is the one between 80 and 90 Hz, which
corresponds to the mode found with the vertical excitation. The resonance
ANALYSIS OF TIRE PROPERTIES 763

•AOlAL-HT TTHC

-
r
'MO

I*

100 HO 100 Hi
00
- 4
TS 35 JB"Mt

FIGURE 9.4.19. Amplitude ratio and phase delay of calculated turn of vertical deformations
distributed along the circumference with respect to platform motion (adapted scale) com
pared with measured hub response offigure 9. 4. 1 7, after being cleared ofthe lowerfrequency
components arisingfrom the vibration ofthe system considered as a rigid mass-spring system.

occurring at about SO Hz does not correspond to the natural frequency of


the tire rotating as a whole with respect to the wheel fixed in space (cf. sec.
9.4.2, table 9.4.2: 45.5 Hz) but represents the natural rotational vibration
of the wheel with a great part of the tire (about a point approximately
midway between hub and upper rim) with respect to the contact patch.
The peak at about 13 Hz will be due to the natural frequency of the sys
tem, mainly owing to the longitudinal compliance of the suspension.

n/2 OUT n OUT


IN PHASE IN PHASE OF PHASE OF PHASE

O 01
20 3O 40 SO 6O 7O SO 9O IOO
FREQUENCY — Hz
The magnitude and sense of the vibration are shown by the numbers and by the shading of the
arrows. The phase with respect to the input for all the stations on any one diagram is shown above
the diagram. The unity transmission ratio line defines the magnitude of the input to the tyre.
The transmission ratio curve itself refers to longitudinal vibrations at the axle.

FIGURE 9.4.20. Modes of vibration of the suspension-wheel-tire (radial-ply) combination for


longitudinal excitation by a vibrating platform.
The magnitude and sense of the vibration are shown by the numbers and by the shading of the
arrows. The phase with respect to the inputfor all the station on any one diagram is shown above
the diagram. The unity transmission ratio line defines the magnitude of the input to the tire. The
transmission ratio curve itself refers to longitudinal vibrations at the axle.
764 MECHANICS OF PNEUMATIC TIRES

The experiments discussed above are carried out on a tire mounted on a


hub which must be capable of moving in order to obtain the overall trans-
missibility. The natural frequencies of the tire itself might be measured
more accurately on a tire mounted on a rim which is fixed in space. The
results of some of these investigations will be given in the subsequent
paragraph as an illustration of the theory (cf. figs. 9.4.24-25).
Tread-Band Free Vibrations
For the development of a theory for the plane vibrations of a tire one
may consider a circular membrane under tension or a circular beam or a
combination of these two. In the early literature we encounter a number of
theories for the vibration of a two-dimensional elastically-supported circu
lar beam or shell [30, 31]. More recent studies by Tielking [32] and Böhm
[33] and a more simple theory by Fiala and Willumeit [34] lay specific em
phasis on the vibration of a pneumatic tire. The complicated case of a to
roidal shell has been solved, after many simplifying assumptions, by Fed-
erhofer [35]. In the theory presented below we treat the vibrations of
circular beam and make use of the method employed by Tielking, which is
based on the principle of Hamilton. In an example we present the equa
tions and results as given by Böhm, which also cover vibrations in the lat
eral direction.
Unlike Tielking, we introduce a tension force acting in the beam which
may be due to the internal pressure and the rotational speed. We also ac
count for the tangential stiffness between beam (tread-band) and wheel
rim. In figure 9.4.21 the pneumatic tire model is shown. Damping is not
considered.
The application of Hamilton's principle requires the determination of
the potential and the kinetic energy of the model as well as the work done
by internal pressure forces.
The potential energy is composed of a part due to elastic deformations
of the beam and a part due to radial and tangential displacements of the
tread band with respect to the wheel rim. For the determination of these
quantities we introduce the independent angular coordinate θ and the tire
radius r, representing the position of a tread element at rest with respect to

clastic
foundation

FIGURE 9.4.21. Cylindrical beam or shell model for a pneumatic tire.


ANALYSIS OF TIRE PROPERTIES 763
a coordinate system fixed to the wheel. Furthermore we introduce the de
pendent coordinates u = n// and w, denoting the tangential and radial dis
placements, respectively, of the element with respect to the original posi
tion.
Figure 9.4.22 shows an element in the original position (possibly influ
enced by the centrifugal force) and in the deformed and deflected situa
tion. The tangential strain at the original radius rz = r + z is influenced by
the initial strain c» (due to internal pressure and centrifugal force), by the
additional "angular" strain tty/dθ and by the radial displacement w and its
derivatives. The strain becomes:

€, -1
with
r, = r + z.
The potential energy due to tread band deformations becomes, per unit
width:

where the integration is over the tread band thickness h and in which £
denotes the elastic modulus of the beam material. With the introduction of
u = rψ, I = 1/66A3 (moment of inertia) and S0 = cje, (tension force with cs
denoting the tensile rigidity) we obtain, with the assumption that e. and
(h/r)2 are negligible with respect to unity:
?w dwV
w (9.4.83)
dff2 del
in which only the second-order terms are shown.
The potential energy stored in the elastic foundation becomes, using cr
and ct to denote the radial and tangential stiffness per unit length:

(9.4.84)

The work done by the internal pressure p, is:

with

and
r + w.
766 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4.22. Displacement and deformation of tread band element.

When retaining only the second-order terms we obtain:

W, = bp, 2w + w> dO. (9.4.85)

For calculation of the kinetic energy we consider figure 9.4.23. The wheel
rotates with an angular velocity £2. In addition, a tread element has a ve
locity with respect to the wheel. The velocity vector of such an element ex
pressed in terms of the unit vectors in radial and tangential direction e,
and e, becomes:

With the use of this expression we obtain the following formula for the ki

prde

FIGURE 9.4.23. Position of an element and its velocity.


ANALYSIS OF TIRE PROPERTIES 767

netic energy of the shell per unit width with p denoting the mass per unit
area:

•/ {w* + (fl + $)\r + w)2} d6


'0

which becomes when retaining only second-order terms and with m = r$r.

T - Vipr I {w* + SPw2 + i? + ASlwu) dO. (9.4.86)

According to Hamilton's principle the time integral of the Lagrangian L


must be minimized, so that:
L=T-V+Wr. (9.4.87)
The Lagrangian density dL/dd is easily formulated by combination of the
above expressions. We obtain:

r \d0
du du
(9.4.88)

The time integral of this quantity becomes stationary when the following
Euler-Lagrange differential equations for the dependent variables w and u,
which hold for our case of two independent variables θ and t, are satisfied
[36]. With the abbreviations ( )′ = d( )/M and (') - 3( )/dt we obtain
dL' a dL' a dL' d2 dL'\
-o,
dw ' dt dw M dw' dff2 dw"
dL
' r a dL'
, 3 j dL'\ — n (9.4.89)
di < ft du 1 a e do1

Application of these equations gives the following equations of motion for


the cylindrical beam of unit width:

p(w - - (wiv + 2w" + w) - w"

c.
(u' + w) + c,w = 0, (9.4.90)

c,u - -j nO + — w'-O. (9.4.91)

Working along similar lines Tielking obtained almost the same equations.
768 MECHANICS OF PNEUMATIC TIRES

However in the dynamic part of the second equation the additional term
—pfl2u appears in his theory, while Tielking did not consider S0 and c,.
The remaining analysis will be simplified by using the concept of in-
extensibility of the cylindrical beam (c, -* oo). It is believed that this ap
proximation is particularly good when the consideration is restricted to
radial ply belted tires. Mathematically this may be accomplished by put
ting:
u7 - -w. (9.4.92)
We obtain by elimination of the terms with coefficient cs in the equations
(9.4.90) and (9.4.91):
El
-w + w")

W" - c,w = 0. (9.4.93)


\
In an analogous way Bryan [31] derived essentially the same equation.
In his study of oscillatory motions it is assumed that the radial dis
placement w varies harmonically in the form:
w = A sin (s6 + wt). (9.4.94)
Substitution in eq (9.4.93) yields

S2V) + (-J6 + 2j4 - s

(9.4.95)

with the definition

^c,\(s2+\) (9.4.96)

we obtain
2Sfc
w = -:—r ± w. (9.4.97)

The possible solutions become herewith:

>V| = A , sin ut
J2+1
(9.4.98)
™ A2 sin / - ut\ .
ANALYSIS OF TIRE PROPERTIES 769

In the particular case A 1 = A2 = A we obtain for the solution w = w1 + w2:

2A sin cos ut (9.4.99)

in which to obviously represents the frequency of the vibration. The mode


number ,v denotes the number of periods along the tire circumference. The
number of nodes is twice as much. The position of a node

changes with respect to the wheel with an angular speed fi*:

a (9.4.101)

For the observer stationary in space the nodes move with a different veloc
ity, fi*, which obviously equals 12* + 12:

2+ 1 (9.4.102)

For each number of periods s the frequency of the vibration may be calcu
lated with the aid of equation (9.4.96). The frequencies depend on tire pa
rameters and to a very small extent on the rotational speed Ω. The tension
force S,, may be a function of £2 (cf. sec. 9.4.2 on standing waves).
Böhm, who did not consider a rotating tire, calculated the values of
these parameters from the tire construction, geometry and material prop
erties. Böhm's equations differ in some respects from those derived above
[90, 91]. Böhm's equations, for the derivation of which we refer to the orig
inal article [33], read:
P* + ^T Kv + *"") + fr (tf + w) - ^r (W + w) + crw - 0, (9.4.103)

pit + ££- (w"' + wO - ^ (u" + wO -I- CM - 0. (9.4.104)

With the assumed solutions


w = A sin wt sin s6, u -= B sin wt cos s6 (9.4. 105)

TABLE 9.4.2 Natural frequencies of symmetric (plane) tread band vibrations


s"0 s*-l s~2 j=3 s-4 s-5
Measured n(Hz) 83 98.5 115 136 158
Calculated n(Hz)
(eqs(9.4.103)) 45.5 83.7 105.5 119 134 150
Calculated n(Hz)
(eq (9.4.96)) 45.5 87 109 123 138 154
770 MECHANICS OF PNEUMATIC TIRES

inserted in these equations the natural frequencies can be obtained for the
various mode numbers .v. The calculations were carried out by Böhm for a
135-13 radial ply Michel in X tire with inflation pressure pi = 1.25 bar. The
following calculated values hold for the total width 2b(= 0.075 m.): ρ =
1.25 kg/m., EI = 0.7 Nm.2, cs = 59 × 104N, S0 = 1920 N, cr = 75 × 104 N/
m.2, ct = 12.8 × 104 N/m.2, r = 0.273 m.
In table 9.4.2 both the calculated and measured frequencies n = ω/2ir
(Hz) are shown for mode numbers 0-5. For comparison we have added
the results obtained with the equation earlier derived (96) for 0 = 0, so
that ω = w. For each mode number a second natural frequency exists
which, according to the theory, is 4 to 8 times higher than the lower one
shown in table 9.4.2. The frequency 45.5 Hz at mode number s = 0 corre
sponds to the mode shape showing pure tangential torsion of the tread
band with respect to the wheel rim. The second frequency at s = 0, which
occurs at 376 Hz exhibits purely uniform radial displacements of the tread
band. The experiments were carried out by applying a radial excitation
force so that the low frequency zero-mode natural frequency was not de
tected.
In figure 9.4.24 the theoretical and measured mode shapes are pre
sented. The experimental frequency response characteristic for the radial
displacement amplitude A, with respect to the radially applied excitation
force amplitude (constant) has also been indicated. The correspondence
between theory and experiment is good enough for technical purposes.
The differences between eqs (9.4.90-91) and (9.4.103-104) turn out to be
of no importance in the range of mode numbers investigated. Also, it ap
pears that the concept of inextensibility used in eq (9.4.93) is acceptable
for this tire. From the calculations it appears that the influence of the
bending stiffness EI starts at higher values of s. At s = 5 the influence of
EI on ω amounts approximately 2 percent. We shall see that in the study
of standing waves with relatively small wave lengths the influence of the
bending stiffness is appreciable.
Lateral vibrations
For the same 135-15 Michelin X tire Böhm also studied the lateral vi
brations. We shall restrict ourselves here to a short presentation of his the
oretical and experimental results. For the derivation of the equations we
refer to the original paper [33].
Again two degrees of freedom are considered for an element of the tire
tread-band. The variables are: the lateral deflection c and the torsion angle
/? about a tangent to the peripheral line. These so-called anti-symmetric
variables are not coupled with the symmetric variables u and w. With the
assumed solution
v = A sin ut sin s6 and fi = B sin ut cos sff (9.4.106)
the following eigenvalue determinant is obtained:

12
(9.4.107)
ANALYSIS OF TIRE PROPERTIES 771

THEOR Y

EXPERIMENT

FIGURE 9.4.24. Theoretical and experimental mode shapes and measuredfrequency response
characteristic of the radial tread band deflections with respect to a radial excitation force for
a radial-ply Michelin X 135-13 tire with p, - 7.25 bar (from Bdhm [33]).
772 MECHANICS OF PNEUMATIC TIRES

TABLE 9.4.3 Natural frequencies of anti-symmetric (lateral) tread band vibrations


j-0 s-2 j-3
Measured n(Hz) 39 44 76 114 149
Calculated n(Hz) 40.0 42.7 64.2 110 162

in which S, = 1920 N (total tension force), GIp = 3.2 Nm.2 (torsional stiff
ness of tread band), EIz = 150 Nm.2 (bending stiffness of tread-band about
wheel radius), ct = 12.8 × 104 N/m.2 (tangential foundation stiffness), cc =
105 N/m.2 (lateral foundation stiffness), ct = 1050 N.m/rad.m. (torsional
foundation stiffness), ;>, = 1.25 × 105 N/m.2, ρ = 1.56 kg/m., r = 0.273 m.,
2b = 0.075 m. (width of carcass breaker), b* = 0.056 m. (reduced width),
t = 0.003 m. (side wall thickness).
In table 9.4.3 the calculated and measured natural frequencies are pre
sented for mode numbers s = 0-4. Figure 9.4.25 shows the measured mode
shapes of the lateral deflection and in addition the frequency response
characteristics of the lateral tread deflection with respect to an excitation
force acting upon the tread-band in purely lateral direction.

FIGURE 9.4.25. Experimental mode shapes and measured frequency response characteristic
ofthe lateral deflection of the tread band with respect to a lateral excitationforce acting upon
the tread band.
Same tire u in figure 9.4.24 (from Bohm [3)]).
ANALYSIS OF TIRE PROPERTIES 773

Bohm found the mechanical characteristics of the tire by working back


wards from the measured spectra of natural frequencies, and found good
agreement with the above values which were calculated from tire dimen
sions and material properties. Böhm states that the backward calculation
of physical values is a powerful method, since the measured natural fre
quencies describe the system more precisely than do static deformation ex
periments.

Standing Waves
When rolling at high speed, waves are formed on a tire behind the area
of contact with the road. The repeated deformation caused by the wave
process results in a considerable heat buildup which reduces the strength
of the tire and may lead to its ultimate destruction.
Because these waves present a stationary appearance to an observer
they have been called standing waves. Figure 9.4.26 gives a very good il
lustration of such a stationary wave deformation. Gardner and Worswick
[37] published considerable experimental information on this phenome
non. Turner [38] also provides interesting theoretical and experimental in
formation on amongst other things the power consumption owing to
standing waves. More recently Drozdov, et al. [39], Togo [40], Böhm [33]
and Fiala [34] presented theories on standing waves. Ames [41] prepared a
literature survey in which mention has been made of other Russian and
Japanese work. Ames proposes the introduction of nonlinear elements in
the dynamics of the tread-band and side wall motion. Akasaka and Yam-
agishi [42] studied the standing waves in the shell wall of a running tire by
considering the tire as a cord reinforced toroidal membrane shell with el
liptical cross section. The cords have been assumed to be inextensible, the
flexural rigidity has been neglected, and tangential displacements have
been neglected. Solutions have been obtained by means of Galerkin's
method. Recently, Soedel [45] has analysed dynamic tire behaviour using
dynamic Green (influence) functions formulated in terms of natural tire
frequencies and modes. A complete solution of the dynamic response of a
rolling inflated tire has been given, in principle, for three-directional load
ing conditions. It was shown that during steady-state constant-speed roll
ing all tire responses are stationary (standing waves). Under the assump
tion that a concentrated normal force acts upon the tire and remains
constant (independent of time and speed) it has been found that in the un
damped case critical situations are expected when the angular rolling
speed coincides with one of the natural frequencies divided by its circum
ferential wave number.
. In the theories enunciated below the finite length of the tire circum
ference has not been taken into account, in view of the practical observa
tion that standing waves occur only in a region close to and behind the
contact area. They do not appear to extend around the complete circum
ference but damp out on their way to the front edge of the contact zone. It
is appreciated, however, that Soedel's approach forms a valuable contri
bution to the theory of tire dynamics. Padovan [46] has also numerically
simulated the dynamic behavior of a visco-elastic tire model. He presents
standing waves graphically as functions of speed and tire parameters. In
the theories to he presented below we shall restrict ourselves to shells of a
circular cylindrical shape. Tangential displacements will also be neglected
here.
774 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.4.26. Standing wave formation at high speeds of travel (from B. Nylon S., Great
Britain).

Membrane theory
The equation of the transverse motion of a membrane stretched in lon
gitudinal directions reads:

(9.4.108)
dx1
where ρ represents the mass per unit length, S the tension force, n> the
transverse displacement, x the longitudinal coordinate and t the time. Af
ter a disturbance a wave is formed which propagates with a velocity

aP (9.4.109)

It is expected that when a normal load is applied which travels with re


spect to the membrane with a speed equal to the above propagation veloc
ity a critical situation occurs. Togo [40] has investigated this problem for
an elastic-ally supported membrane and found that for speeds in excess of
this critical value a standing wave is formed behind the point of loading.
In front of this point the displacement vanishes.
We shall examine the problem of a circular membrane or string under
tension, radially supported by an elastic foundation which may show some
viscous damping. The tension is supplied by the inflation pressure. This
model tire rolls over a perfectly smooth horizontal surface and is loaded
vertically. We introduce the quantities; r = tire radius, Ω = rotational
speed; V ≃ flr = speed of travel; W = normal load; pi = inflation pressure;
cr = radial stiffness of foundation per unit length; kr = coefficient of radial
damping per unit length; cs = tensile carcass stiffness of unit length (= av
erage elastic modulus times tread-band cross section); θ = angular coordi
nate with respect to an axes-system fixed to the wheel; w = outward dis
placement of rotating membrane due to W; TJ = percentage of centrifugal
force restored by radial forces; 1 - 17 = percentage of centrifugal force re-
ANALYSIS OF TIRE PROPERTIES 775

^e

FIGURE 9.4.27. Forces acting on membrane element.

stored by tangential tension forces; S0 = tension force in nonrotating tire;


2b = effective tread-band width.
It is assumed that only radial displacements occur. In figure 9.4.27
forces acting on a membrane element of unit width are indicated. The fol
lowing partial differential equation applies for the radial displacements w
(6, t) due to the load W in zones outside of the contact area. Consequently,
the growth of tire radius due to centrifugal forces is not included in w.

(9.4.110)
where
(9.4.111)
This equation corresponds to eq (9.4.90) for u = 0 and EI = 0 but with the
additional tension force due to &22. The solution of this equation represents
the tire deformation seen with respect to a rotating coordinate system. For
the description of the standing wave phenomenon we shall adopt coordi
nates with respect to a system fixed in space. The angle 4> is introduced in
dicating the position of a tread element with respect to the vertical through
the wheel axis (cf. fig. 9.4.27),
(9.4.112)
Equation (9.4.110) then assumes the following form:
d*w . dw

0. (9.4.113)
776 MECHANICS OF PNEUMATIC TIRES

In case of stationary appearance of the radial deformations all derivatives


with respect to time vanish. We obtain the ordinary differential equation:

o (9.4.114)

which could have been obtained immediately from (9.4.110) by putting


d/dt — -Sld/d<t> and d/90 = - d/d<f>, which transformations are valid in
case of deformations being stationary with respect to a coordinate system
fixed in space. The general solution of (9.4.1 14) reads:
(9.4.115)
in which the roots of the characteristic equation become:
s.
*
(9.4.116)

Comparison with existing theories of Böhm [33], Fiala [34], and Togo [40],
the latter two being reduced here to the case without bending stiffness of
the tread-band, reveals that their results deviate from (9.4.116) in the fol
lowing respects. Fiala and Togo neglect the effect of damping (kr = 0),
Togo and Böhm assume that the centrifugal force does not influence the
tension force 5 (77 = 1), Fiala assumes the opposite (TJ = 0) and Togo ne
glects pSl2 with respect to c′.
For small values of kr the solution (9.4.1 16) may be written as:
-kft
(9.4.117)

These roots become complex and consequently the solution (9.4.1 15) be
comes oscillatory when 12 satisfies the following condition
0,<a<Q, (9.4.118)
where 12, and 122 are critical values of rotational speed expressed by: .
"Jo *-*•» C
0?- (9.4.119)
Before discussing the solutions in greater detail, we shall first examine the
order of magnitude of fl, and Ω2, according to a number of known theories.
For a 135-13 Michelin radial ply tire Böhm obtains by experimental
means the following physical values: S0 = 1920 N., p- 1.56 kg/m., cs = 59
× 104 N., cr = 75 X 104 N/m.2, 1pt> ≃ 104 N/m. (pi = 1.25 bar), r = 0.273
m.
We obtain with ij = 1:
0, - 128*- (K, - 126 km/hr.),
S22 = 2350*- ' (F2 - 2300 km/hr.).
For a Bridgestone 165-400 4PR diagonal ply tire investigated by Togo (to
ANALYSIS OF TIRE PROPERTIES 777

be treated later on in greater detail) the following theoretical values can be


derived c′/2b = 5 × 107 N/m.3 (≃ 5 kgf/cm.3), SJTfb = 136 X 103 N/m.3
(≃ 0.0136 kgf/cm.3), ρ/2b = 21.8 kg/m.2 (≃ 2.22 X 10-6 kgf · sVcm.3), pi =
13 × 104 N/m.2 (= 1.3 bar), r = 0.34 m., which values result with 77 = 1 in.:

B2 = 1500.T1 (K2 = 1830 km/hr.).


For a 165-400 radial ply tire we obtain, according to the formulation of
Fiala [34], c′ = 35 × 104 N/m.2, ρ = 2.1 kg/m., r = 0.33 m. With Fiala's
assumption 17 = 0 the following results are obtained for pi = 1 bar:

J22 = 410^-' (Vt = 480 km/hr.).


The three theories deviate considerably from each other. Apart from the
differences in TJ we have noted, amongst other things, that Fiala took cs =
0, which, in particular for a radial tire, is believed to be an unacceptable
assumption. It causes the low value of S22 obtained with Fiala's theory.
We continue now the discussion of the solution and write for (9.4.1 17),
with the aid of (9.4.1 19):

|P_Q2- (9.4.120)
Two different classes of values will be considered. The first category refers
to the theories of Togo and Böhm which hold for Ω2 > Q,. We obtain for
the speed ranges:
a. Ω < S2,: Real roots A, and A,,. Assume k, small enough so that for the Ω
considered one positive (A,) and one negative root (A2) exist. With the re
quirements that for |</>| —» oo the deflection w remains finite the solutions
(9.4.1 15) become for the regions outside of the contact area with length 2a:

(9.4.121)
a_
r'
indicating an almost symmetrical monotonically damped deflection. In the
neighborhood of J2,, the deflection at the rear must vanish due to the pres
ence of fc,(λ2 < λ1 < 0, C1 = 0).
b. J2, < J2 < Ω2: Complex roots λ1,2 = α ± //? with α and ft real positive
numbers. With the same conditions (|$| -» oo, w finite) we obtain in this
range of rotational speed:

(9.4.122)
sin (fo + &) = 0, (C2 = 0).
Consequently, in this speed range a standing wave will be formed behind
the contact area. The wave damps out due to the presence of the radial
778 MECHANICS OF PNEUMATIC TIRES

damping coefficient kr. In front of the contact area the solution would in
dicate an exponentially increasing deflection w which is in disagreement
with boundary conditions, so that the constant of integration C2 must van
ish, resulting in a front portion of the tire tread-band of which the deflec
tion w vanishes. Figure 9.4.26 shows a beautiful picture of the actual de
formation when operating in excess of the critical speed.
The theoretical wave length of the standing wave varies with speed ac
cording to the following formula:
2irr
(9.4.123)

Figure 9.4.28 gives a graphical representation of the variation of / with Ω.


It should be noted that the wavelength increases with 0 from zero to in
finity, at which fi, has been reached.
r, i2 > Ω2: This is a fictitious case as the tire will be destroyed long before
this very high speed range is reached. The theory predicts two real roots
for the stationary solution. A stationary situation, however, will never be
reached since with the unloaded undisturbed tire, where derivatives with
respect to θ vanish, according to eq (9.4.1 10) an unstable situation occurs.
The last coefficient becomes negative in this speed range and the tire ex
plodes.
In the second category of values fl, > Ω2 which refers to the theory of
Fiala, we distinguish:
a. il < fij,(<fi,): Two real roots exist resulting again in a practically sym
metric damped deflection on both sides of the contact area.
b. £2, < tt < S2,: When a stationary deflection would have been attained,
two complex roots λ1,2 = - α′ ± //?' arise with α′ and β′ representing posi
tive real numbers. We find now a damped standing wave in front of the
contact area (not recognized by Fiala) and vanishing deflections at the
rear. The wavelength appears to decrease with increasing Q until Ω1 is at
tained, where the wavelength vanishes. This decrease corresponds to theo
retical findings of Fiala who did not neglect the bending stiffness of the
carcass tread band. This, however, does not influence the value of the crit
ical rotational speed Ω2. Also when the bending stiffness is introduced, the

FIGURE 9.4.28. Wavelength I of the membrane type tire model as a function of speed Vfor
the case that ii_, > Q,.
ANALYSIS OF TIRE PROPERTIES 779

tire is theoretically expected to explode in excess of Ω2 so that Fiala's the


ory does not seem to have practical value. The third speed range, Ω > S2,,
is of no interest to our analysis.
Influence offlexural rigidity.
The introduction of bending stiffness gives rise to additional terms in
the equation of motion. Consideration of equation (9.4.90) with u = 0, de
rived in the preceding section, shows the additions to be introduced. In
case of deformations stationary with respect to an axes-system fixed in
space, we may write ∂//∂θ = — d/d<j>. Equation (9.4.114) now becomes,
when damping is omitted:
.El S. M d^ .(.. El „ 0 (94124)
dtf ' [ r* '
This equation differs in some respects (El (w"" + w") instead of El \
+ 2w" + w)) with respect to the equations of other authors (Togo, Böhm,
Drozdov, Fiala). We shall adopt the simplification introduced by Togo,
viz. omission of pfl2 relative to c′, and follow the line of his theory [40].
The following quantities are introduced for the sake of abbreviation:
EI ~ S" n|=C' + ~'" • (9.4.125)

With Togo's simplification we obtain for the solutions ω = exp (A<£) the
following characteristic equation:
r*\4 + (2S - O? + J22)A2 + J2| - 0 (9.4. 1 26)
with four roots determined by:

X2 = — ^ Q — (9.4. 127)

-a± V"2 - (®2/rf-


The discriminant vanishes at two speeds II, and S2, :
: - 2»\
(9.4.128)

These speeds represent the lower and upper critical rotational speeds for
the stretched beam type of model, whereas S2, is the critical rotational
speed for the stretched string type model. The solution (9.4.127) may be
divided into three speed ranges 1, 2 and 3 also indicated in figure 9.4.29:
1. Ω < flt: Four real roots. The deflection curves are combinations of two
damped exponential curves.
2. QL < ft < ΩU: Four complex roots. The deflection curves are of a
damped oscillatory character.
3. Qy < $2: Four imaginary roots. Standing waves are formed composed of
two different modes.
As an effect of the bending stiffness, a range of rotational speed arises in
780 MECHANICS OF PNEUMATIC TIRES

Im

standing txpo -
waves damped nentially
(two modes) standing wave damped

IJTA n Q.V=0

FIGURE 9.4.29. Solutions of\2 according to eq 9.4.127.

which damped waves are formed in front of and behind the contact area.
This speed range S2, < Ω < S2,, is situated on both sides of the value
V(fl2 — v\ which represents the critical rotational speed arising when in
eq. (9.4.124) the fourth derivative is omitted.
Above the upper critical speed B^ the standing wave appears only be
hind the loading point. At fi = S2,; the standing wave already has a finite
wavelength. Beyond this critical speed two wavelengths arise, one of
which increases and the other decreases with increasing Ω, as may be de
duced from figure 9.4.29. Turner [38] reported that such a combination of
wavelengths indeed can occur in practice. It is expected, however, that in
most cases the shorter waves are suppressed as a result of material damp
ing.
Example of calculation of the critical speed and wavelength for a bias ply
tire.
The following equation which corresponds to eq (9.4.124) seems to have
been used by Drozdov and others [39] in their calculations:

D-B A- (9.4.129)

The expressions for parameters A, B, C and D used by Drozdov may be


found in the original paper. The quantity r1 denotes the radius of the tire
cross section (half tire width) and £ = x/r1 represents the dimensionless
tangential coordinate (x = <f>r).
Togo [40] adopted the same equations but used different expressions for
A, B, C and D. We shall present here the theoretical and experimental re
ANALYSIS OF TIRE PROPERTIES 781

suits of Togo for a bias ply Bridgestone tire with the following data (1 kgf.
≃ 10 N): Tire 165 × 400 4 Pr SAD; outer diameter 2r = 68 cm.; tire width
2r1 = 17 cm.; inflation pressure pi = 1.3 kgf/cm.2, tire load 300 kgf.
The tire characteristic values to be used in the analysis are: crown angle
= average cord angle measured from the wheel center plane α = 40°;
number of cord plies n = 4; number of cords at tire crown per ply and per
cm. width i = 9 cm.-1; average thickness of tread band h = h1 + h2 = 1.9
cm.; distance from neutral ply to outer surface h1 = 1.7 cm.; distance from
neutral ply to inner surface h2 = 0.2 cm.; elasticity modulus of rubber ER =
1 17 kgf/cm.2; Poisson's ratio of rubber ju = 0.2 (0.5 according to others);
mass of tire per unit area p = 2.22 × 10-6 kgf- s2/cm.3; tensile elasticity
modulus of a cord EC = 50 kgf.; flexural rigidity of the ply layers per unit
width AC = iEc X y2 cos4 α = 7.75 kgf- cm. (y = distance of ply from
M

neutral ply); flexural rigidity of rubber tread-band AR = Vi ER (h\ + hi)/(I


— /i2) = 200 kgf · cm.; it has been assumed furthermore that the centrifugal
force does not influence the tension S of this bias ply tire so that TJ = 1.
From Togo's analysis it can be deduced that the dimensionless parame
ters A, B, C and D of eq (9.4.129) are expressed as follows (the relations
with parameters of eq (9.4.124) are given in between brackets):

ERhr\\ + tan4 a - 2/i tan2 a)

-2EI Ac + A,
B cot2 a —
Pf\ (9.4.130)
EIr2 \
C
r\

Substitution of the solution w = e?K, β real, yields relations which hold in


the speed range Ω > ΩU where standing waves are formed behind the con
tact area. The dimensionless reduced frequency ft is inversely proportional
to the wavelength /,

(9.4.131)
We obtain the equation:

C/T-i i-B A- 0 (9.4.132)


\Pf Pf^

with the solution:

D-B± >-B -4C A- (9.4.133)

Again, two wavelengths occur theoretically. When the discriminant van


ishes, the critical value of speed V., is attained above which si
782 MECHANICS OF PNEUMATIC TIRES

1 ut ions of w exist. At this point of transition the values of V, ft and l will be


provided with the index U. We derive:

(9.4.134)
I*"" /
in which

fi>

Expression (9.4.135) is the same as the one used by Drozdov. With the tire
data listed above we obtain: VU = 201 kmhr. and lU = 16 cm. With eqs
(9.4.127-128) we find practically the same result for V^= rΩU) and /„, the
lower critical speed VL becomes imaginary. For V1 = r Ω1, which is the crit
ical speed in case of the absence of bending stiffness, we obtain: K, = 96
km/hr. Considering equation (9.4.124) once again, it may be concluded
that Togo's simplification (pfi- <c c') and the omission in the coefficients of
w" and w of El/i* (Bflhm, Fiala, Togo, Drozdov) are permissible since
their effects appear to be extremely small.
Togo has carried out a number of test runs on a steel drum using the tire
described above. Wavelength and amplitude of the standing wave have
been measured from photographs. Figure 9.4.30 shows the measured rela
tion of wavelength versus speed. At the speed 140 km/hr., Togo did not
observe a wavy deformation but the endurance test was ended after four
minutes due to ply separation.
Comparison with theoretical results suggests that the theory only pro
vides an approximate insight into the problem. Drozdov reported similar
discrepancies between theory and experiment. Attempts have been made
to obtain a better agreement through the introduction of an effective
(smaller) tread thickness for the calculation of the flexural rigidity. This
may be supported by figure 9.4.26 (rear view) which clearly shows the
manner in which the tread rubber deforms. If we assume that the effective
thickness is half of the geometric thickness h, then the quantity Ac + AR
becomes approximately one-eighth of its originally calculated value. The
critical values become now: Vu~ 150 km/hr. and lU = 9.5 cm., which re
sults are closer to the experimental values of figure 9.4.30.
From the analysis above and in particular from eq (9.4.134) with eq
(9.4.130) it may be concluded that the critical speed is shifted to higher
values when:
1. bending stiffness of tread EI is enlarged (n enlarged, a reduced),
2. tension force in tread S.. is enlarged (pi enlarged),
3. radial tire stiffness cr is enlarged (shape cross section, sec. 9.4.1).
4. tensile tread-band rigidity cs is enlarged (a reduced),
5. percentage of the centrifugal force restored by radial forces TJ is re
duced (α reduced),
6. mass density ρ is reduced.
ANALYSIS OF TIRE PROPERTIES 783

Jt
(cm)
X
0 40 80 120 160 200
V (km/hi

FIGURE 9.4.30. Experimentally obtained relation between wavelength I of standing wave and
speed Vfor a Bridgestone cross-ply tire 165-400 (p, -1.3 bar) on a steel drum. (Togo [40]).

The influences of inflation pressure pi, crown angle a and number of plies
n are in accordance with experiments of Gardner and Worswick [37] (bias
ply). Experiments of Curtiss show that unfavorable interaction between
parameters may occur. From measured rolling resistance curves (cf. fig.
8.2.60) it is seen that radial and bias ply tires behave differently when the
mass of the tread-band is reduced through removal of the tread. As ex
pected, the bias ply tire then shows an increase in critical speed above
which standing waves occur, causing a further progressive rise of the roll
ing resistance with speed. For the radial tire, however, it appears that the
critical speed decreases, which, according to Curtiss, may be explained by
the loss in carcass sidewall rigidity due to a reduction in tension in the
sidewall caused by the decrease in tread-imposed centrifugal force.
References
[1] Hadekel, R., The mechanical characteristics of pneumatic tires, S & T Memo, No. 10/
52 (British Ministry of Supply, TPA 3/TIB, 1952).
[2] Weber, G., Theorie des Reifens mit ihrer Auswirkung auf die Praxis bei hohen Beans-
pruchungen. A.T.Z. 56 (12), 325 (1954).
[3] Marquard, E., Untersuchung iiber den Einfluss der Stossdämpfer auf die Zwischen Rad
und Fahrbahn auftredenden senkrechten dynamischen Bodenkräfte. Deutsche Kraft-
fahrtforschung 104, 32 (1957).
[4] Chiesa, A., and Tangorra, G., The dynamic stiffness of tires, Revue Gin. du
Caoutchoue 36(10), 1321 (1959).
[5] Rasmussen, R. E., and Cortese, A. D , Dynamic spring rate performance of rolling tires,
SAE Paper No. 680408 (1968).
[6] Betz, E. R., Studying structure dynamics with the Cadillac road simulator, SAE Paper
No. 660101 (1966).
[7] Dodge, R. N., Dynamic stiffness of a pneumatic tire model, SAE Paper No. 65049 1
(1%5).
[8] Rotta, J., Zur Statik des Luftreifens, Ing. Archiv 1949, p. 129.
[9] Senger, G., Ueber dynamische Radlasten beim Ueberrollen Kurzwelliger Unebenheiten
durch schwere Luftreifen, Deutsche Kraftfahrtforschung 187 (1967).
[10] Pacejka, H. B., Theoretische beschouwingen over het gedrag van luchtveren, De Inge-
nieur 72(10), (maart 1960).
[11] Clark, S. K.., An anlytical model for lateral stiffness of a pneumatic tire, Paper 3-01,
FISITA, Barcelona, 1968.
[12] Tiemann, R., Aehnlichkeitsbeziehungen im Federungsverhalten von Luftreifen, A.T.Z.
65(4), 97 (1963).
784 MECHANICS OF PNEUMATIC TIRES

[13] Smiley, R. F., and Home, W. B , Mechanical properties of pneumatic tires with special
reference to modern aircraft tires, NASA Tech. Note 41 10 (1958).
[14] Cough, V. E., Tires and Air Suspension, Advances in Automobile Engineering, Tid-
bury, ed., (Pergamon Press, 1963).
[IS] Julien, M. M. A., and Paulsen. M. J., Méthode expérimentale de mesure et definition du
pouvoir absorbant du pneumatique, J. de la S.I.A., janvier 1953, p. 33.
[16] Lippmann, S. A., Piccin, W. A., and Baker, T. P., Enveloping characteristics of truck
tires, SAE Trans. 74(1966).
[17] Lippmann, S. A., and Nanny, J. D., A quantitative analysis of the enveloping forces of
passenger tires, SAE Paper No. 670174 (1967).
[18] Barson, C. W., James, D. H , and Morcombe, A. W., Some aspects of tire and vehicle
vibration testing, Proc. IME 182, 3B, 32 (1967-68).
[19] Guslitzer, R. I.., On the interaction of a motor-vehicle tire with an obstacle, Translation
817, Rubber Res. Ass. British Rubber Manufactures, 1957.
[20] Hey, K. F., Untersuchungen von Längskräften zwischen Reifen und Fahrbahn beim
Ueberfahren von Hindernissen, Diss. T. H. Braunschweig (1963).
[21] Yu Chen, Studies of the interfacial phenomena during braking, Ford Motor Tech. Re
port A-788, 30 (1958).
[22] Hörz, E., Der Einfluss von Bremskraftreglern auf die Brems- und Führungskraft eines
gummibereiften Fahrzeugrads, Deutsche Kraftfahrtforschung 195 (1968).
[23] Holmes, K. E., and Stone, R. D., Tire forces as functions of cornering and braking slip
on wet road surfaces, Proc. Symp. on Handling of Vehicles under Emergency Condi
tions, IME, Auto. Div. (1969).
[24] Reynolds, O , On rolling friction, Phil. Trans. Roy. Soc. 166, 155 (1876).
[25] Fromm, H., Berechnung des Schlupfes beim Rollen deform icrbarer Scheiben, ZAMM
7, 27 (1927).
[26] Julien, M., L'envirage et la tenue de route, S.I.A. (April 1937).
[27] Chiesa, A., Obert, I... and Tamburini, 1... High-frequency vibrations in tires under verti
cal perturbation and their transmission to the wheels, Auto. Engr., 520 (1964).
[28] Chiesa, A., Vibrational performance differences between tires with cross-biased plies
and radial plies, SAE Paper No. 990B (1965).
[29] Barson, C. W., Cough, V. E., Hutchinson, J. C., and James, D. H., Tire and vehicle vi
bration, Proc. IME, Auto. Div., 1964-65, p. 213.
[30] Strutt, J. W. (Lord Rayleigh), The Theory of Sound, Vol. 1, London, 1878, p. 332.
[31] Bryan, J. W., On the beats in the vibrations of a revolving cylinder or bell, Proc Cambr
Phil. Soc. 7, 101 (1890).
[32] Tielking, J. T., Plane vibration characteristics of a pneumatic tire model, SAE Paper
No. 650492 (1965).
[33] Böhm, F., Mechanik des Giirtelreifens, Ing. Archiv 35, 82 (1966).
[34] Fiala, E., and Willomett, H. P., Radiale Schwingungen von Gürtel-Radialreifen, A.T.Z .
68(2), 33(1966).
[35] Federhofer, K., Zur Schwingzahlberechnung des diinnwandigen Hohlenreifens, Ing.
Arehiv 10-11, 125 (1939-1940).
[36] Fliigge, W., Handbook of Engineering Mechanics (McGraw-Hill, 1962), pp. 16-2, 24-6.
[37] Gardner, E. R., and Worswick, T., Behaviour of tyres at high speed, Trans. I.R.I. 27,
127(1951).
[38] Turner, D. M., Wave phenomena in tyres at high speed, Third Rubber Tech. Conf.
London, Trans. I.R.I. (June 1954).
[39] Drozdov, V. K., e.a., Formation of stationary waves on pneumatic tires at high rolling
speeds, Sov. Rubber Tech. 19(12), 36 (1960).
[40] Togo, K., Standing Wave on Pneumatic Tire at High Speed, Memoirs of the Defense
Academy, Japan, Vol. IV, No. 1 (1964), p. 43.
[41] Ames, W. F., Wave phenomena in tires, University of Iowa, Tech. Report 1 (1967),
[42] Akasaka, T., and Yamagishi, K., On the standing waves in the shell wall of running
tyre, Trans. Japan Soc. Aer. Space St. 11(18), 12 (1968).
[43] Clark, S. K., The rolling tire under load, SAE Paper No. 650493 (1965).
[44] Pacejka, H. B., Some recent investigations into dynamics and frictional behavior of
pneumatic tires. Proc. G. M. Symp. Physics of Tire Traction, eds. D. F. Hays and A.
L. Browne. Plenum Press, New York 1974, p. 257.
[45] Soedel, W., On the dynamic response of rolling tires according to thin shell approxima
tions. J. of Sound and Vibration 41 (1975) 2. p. 223.
[46] Padovan, J., On Viscoelasticity and standing waves in tires, Tire Sc. and Technology,
Vol. 4, No. 4, (Nov. 1976), p. 233.
[47] Pacejka, H. B., Berg, J. v.d., and Jillesma, P. J., Front wheel vibrations, VSD-IUTAM
symposium on Dynamics of Vehicles on Roads and Tracks, Vienna, Sept 1977.
9.5. Yaw and Camber Analysis
The so-called antisymmetric (cornering and camber) behavior of tires is
of importance for the investigation of the maneuverability and stability of
automobiles. In particular, the stationary (steady state) characteristics are
employed in such investigations. However, the knowledge of nonsteady
state tire properties is necessary for the investigation of transient motions
or of parasitic motions such as shimmy.
The literature provides a vast amount of experimental data, which in
particular covers the stationary cornering properties of tires. Cornering
characteristics of automobile tires are published, amongst others, by Joy
and Hartley [1],1 Gauss and Wolff [2], Fonda [3], Freudenstein [4] (truck
tires), Nordeen and Cortese [5], and Fonda and Radt [7]. Furthermore, we
mention the more recent experimental studies performed on either real
road surfaces or on drum test stands: Krempel [73], Gengenbach [77], We
ber [69] (inside drum, dry, wet, icy respectively) Henker [6] (inside and
outside drum), Holmes and Stone [74] (wet roads), Fancher e.a. [75, Vol.
1] (wet roads) and Segel [76] (dry roads). These investigations give full ac
count of the interaction between lateral and longitudinal slip. With special
reference to aircraft tires, Smiley and Home [8] have presented a system
atic survey of mechanical tire properties. Hadekel [9] (1952) and Smiley
[10] (1956) gave critical outlines and extensions of existing theories for tire
motions and wheel shimmy. In a recent study, Ho and Hall [67] provide a
vast amount of experimental frequency response characteristics for a set of
aircraft tires measured on a rotating drum together with a correlation
study of theoretical model response curves.
In the theory of tire mechanics relevant to the type of motion we are
considering in this section, the road is assumed to be a smooth level
boundary surface of an undeformable half space, while the tire is repre
sented by some elastic model. The literature provides tire models of vari
ous degrees of complexity. For all these methods the following fundamen
tal observations are applicable.
When the tire moves over the road, horizontal deformations will gener
ally occur over and above the deformation due to static vertical load.
When the wheel moves in such a way that the contact points of an imagi
nary tire, which differs from the real tire only in that it does not show hori
zontal deformations, do not move with respect to the road, we speak of
pure rolling. When all the contact points of that imaginary tire show the
same relative velocity with respect to the road, we speak of longitudinal
(fore and aft) slip or creep when this relative velocity and the rolling ve
locity have the same directions. We speak of lateral (side) slip or drift
when the relative velocity is directed perpendicular to the rolling velocity.
The angle between wheel center plane (direction of rolling) and the vector
of the velocity of the wheel center is called the slip angle. When the wheel
rotates about a vertical axis through the wheel center without showing
longitudinal or lateral slip, we speak of pure spin. When the wheel plane is

' Figures in brackets indicate the literature references at the end of this section.

785
786 MECHANICS OF PNEUMATIC TIRES

tilted with respect to the vertical plane, the tire is said to show a camber or
inclination angle.
A real tire will show additional horizontal deformation. In the case of
dry-frictional contact, the additional horizontal deformations may cause
regions of adhesion as well as regions of sliding. In the following, the
terms slip and sliding will always be used in the sense as expressed above.
In this introduction mention must be made of the work which has been
performed on the problem of the steady state rolling and slipping motion
of two homogeneous bodies pressed to each other. In most of these theo
ries equal elastic constants of the two bodies in contact are assumed (steel-
on-steel problem). The two-dimensional theories of Carter and of Fromm
were followed by three-dimensional theories of Johnson, De Pater, Kalker
and Nayak and Paul (see [19, 21, and 22] for references).
General differential equations of a rolling and slipping body
Consider a rotationally symmetric elastic body representing a wheel and
tire rolling over a smooth horizontal surface representing the road. Fixed
to the road a coordinate system (0, x,y,z)is assumed, of which the x- and
/-axes lie in the road surface and the z-axis points downwards (see fig.
9.5.1). Another coordinate system (C, x, y, z) is introduced of which the
axes x and y lie in the (x, 0, y) plane and z points downwards. The system
moves with respect to the fixed system in such a way that the x-axis lies in
the wheel center plane and the j-axis forms the vertical projection of the
wheel axis. The body (tire) is deformed vertically so that a finite contact
area is present. The center C travels with a constant speed V over the (x,
0, y) plane. The distance traveled s equals
s=Vt (9.5.1)
where t denotes the time. The tangent to the orbit of C makes an angle ft
with the fixed i-axis. With respect to this tangent the x-axis is rotated with
an angle α, defined as the slip angle. The angular deviation of the wheel
plane with respect to the x-axis (yaw angle) is denoted by
t = ft + a. (9.5.2)
For small values of ft the following relation with y. the lateral dis
placement of C, holds
*-£• <9-5-3>
The horizontal displacements of a contact point with respect to its position
in the horizontal undeformed situation with coordinates (x, y) are in
dicated by u and ν in x- and y- direction respectively. The displacements
are functions of x, y and the independent variable s or t.
The position in space of a material point of the rolling and slipping
body in contact with the road (cf. fig. 9.5.1) is indicated by the vector
f-i+q
where s indicates the position of the contact center C in space and q the
position of the material point with respect to the moving system (x, C, vK
ANALYSIS OF TIRE PROPERTIES 787

tro/Klon d v,hwi »,.,

Hi ^M

point on tyre face

FIGURE 9.5.1. Top-view of contact area snowing position with respect to coordinate system
fixed in space (x, 0, y) and deformations (u,v) with respect to moving system (x, C, y).

with e, and e, representing the unit vectors. The vector of the sliding ve
locity of the material point relative to the road becomes (x = dx/dt, etc.):
(x u)e, - (y +
in which V denotes the vector of the speed of travel of the contact center
C. We introduce furthermore:

denoting the vector of the rolling velocity and

representing the vector of the slip or creep velocity of the body with re
spect to the road. We realize furthermore that

l> ~_ ~di_ = Hx ~di_ ^_


~dy~di d7'^

_ Jv dv dx_ dv^dy^ dv
V ~ dt " dx dt By dt dt
788 MECHANICS OF PNEUMATIC TIRES

The sliding velocity becomes herewith in vectorial form:

Henceforth we shall neglect the terms ν · dty/dt and u · dfy/dt, as these are
small of the second order.
The components in x- and y-direction of the sliding velocity Vs of a
point of a rolling body in the contact area with respect to the road are in
general:
. du du du
V" = V" — *"'
wj — V
V" dx V" —
dy -4- —
dt'
(9.5.4)
to ry dy dt

where (Vcx, Vcy) denotes the vector of the creep or slip velocity of the tire,
which is the sliding velocity of the point C of the horizontally undeformed
imaginary tire, which coincides with the center C at the instant considered;
(Vrx, Vry) is the vector of the rolling velocity with which point C moves rel
ative to C. Moreover, ωz denotes the angular velocity of the system (C, x,
y, z) about the z-axis (yaw velocity):

dt dt dt

For better understanding, figure 9.5.1 illustrates the way in which Vv


arises for the case Vry = 0. We will restrict ourselves to small values of lat
eral slip and assume |α| <sc 1. For the system under consideration, i.e., the
tire, where only rolling in the x-direction occurs, the following relations
hold:
V,,= KandF^-0, (9.5.6)
when in addition the longitudinal creep velocity is taken equal to zero,
which is approximately the case when no driving or braking forces are ap
plied. We obtain for the creep or slip velocities:
Vv. = 0 and Vcy = - Va. (9.5.7)

We introduce the variable <f> denoting the spin:

The latter part of this relation holds owing to eqs (9.5.1) and (9.5.5). We
finally obtain the following expressions for the sliding velocities of a point
ANALYSIS OF TIRE PROPERTIES 789

with coordinates (x, y):


du du
^/K=-^>--+-
(9.5.9)

When the vector of the pressure exerted in the positive direction by the
tire upon the road is denoted by (px, />,., pz), we obtain the following rela
tions for the case with finite friction coefficient ft. In the adhesion region,
defined as the area where no sliding occurs ( Va = Vv = 0), the relations
dudu dv dv
dl-d7 = -a + **'
(9.5.10)

hold, and in a sliding region the relations (9.5.9). For the pressure we ob
tain in vectorial form:
(pa pf) = fip,( VM y,,)/ V, (9.5.11)
where
V,= V^+ ^ (9.5.12)
the velocity components VM Vv being determined by (9.5.9).
For the case where only lateral slip occurs (<f> = 0), and, in addition, px =
Vsx = 0 throughout the contact area (which may occur with simplified sys
tems to be treated later on), the relations (9.5.9-11) reduce to:
dv _ dv
djc ds ~ "'
in an adhesion region, (9.5.13)
\P*\

dv dv
in a sliding region. (9.5. 14)
\py\ , sgn*
The equations above apply in general. Their solutions contain constants of
integration which depend on the construction of the tire of which an ap
proximate physical description may be given. In case of a steady state mo
tion, the partial derivatives with respect to the distance traveled ,v become
zero (dv/dJ = du/ds - 0).
Tire models
Many theories are known which describe the qualitative or quantitative
behavior of the steady state or nonsteady state drifting tire. The influence
signifies "sign of."
790 MECHANICS OF PNEUMATIC TIRES

of camber has been studied only to a very limited extent. Due to the im
mense complexity of the tire structure most theories are restricted to a
qualitative description of tire behavior. Particularly in case of the appli
cation of more advanced tire models, this qualitative picture can be
adapted for quantitative use through fitting of the parameters. This can be
done by means of full scale tire experiments, either static or semi-static [4,
1 1, 23] or with the rolling tire [13]. The first method leads to a greater in
sight into the problem, while the second achieves higher accuracy since in
that case the cornering characteristics which are to be fitted are measured
directly.
It should be pointed out that in order to avoid conceptual errors, which
may arise due to the use of oversimplified models, the development of ad
vanced tire models firmly based on actual tire geometry and material
properties is of great importance. In this connection the work of Frank
[11, 12] should be mentioned.
A fundamental difference in structure is apparent between the tire mod
els employed in steady state and nonsteady state tire theories. In most
steady state theories, a model is used consisting of an elastic structure (the
carcass) provided with a great number of elastic tread elements (see fig.
9.5.2). The tread elements contact the road surface in the contact area,
where a region of sliding may occur when the adhesion limit is locally ex
ceeded. In most nonsteady state studies tread elements have been omitted
and, in addition, adhesion has been considered complete in the entire con
tact area. The use of an elastic continuous structure representing the car
cass is essential in nonsteady state tire theories.

9.5.1. Steady State Motion


Side Slip
Models of the carcass commonly encountered in the tire literature can
either be of the beam type or of the stretched string type. The exact repre
sentation of the carcass by a beam instead of a stretched string is more dif
ficult because of the fact that the differential equation for the shape of the
deformed peripheral line of the carcass becomes of the fourth instead of
the second order. For the study of steady state tire behavior, most authors

wheel plane
carcass**
elastic
foundation
tread
rubber ~^^M ""adhesion

y
sliding
r
y
FIGURE 9.5.2. Top-view of tire model in steady state rolling with constant slip angle a.
ANALYSIS OF TIRE PROPERTIES 791

therefore approximate the exact expressions for the lateral deformation of


the beam.
As an extension of the model of Fromm and of Julien (cf. [9] for refer
ences) who did not consider carcass elasticity, Fiala [13] and Freudenstein
[4] developed theories in which the carcass deformation has been approxi
mated with a symmetric parabola determined only by the lateral force.
Böhm [15] and Borgmann [16], the latter without tread elements, use
asymmetric approximate shapes determined by both the lateral force and
the aligning torque. Moreover, Böhm considers in his steady state side slip
theory a two-dimensional contact area provided with a finite number of
tread elements which, due to the width of the tread, will also deform in
longitudinal direction under lateral carcass deformation. As an additional
complexity, Borgmann and Böhm both introduced a coefficient of friction
which is a function of the sliding velocity.
Frank [11] has carried out a thorough comparative investigation of the
various one-dimensional models. He employed a general fourth order dif
ferential equation with which stretched string, beam and stretched beam
tire models can be examined. He obtained the exact solution of the sta
tionary side slip problem with the aid of a special analog computer circuit.
A correlation with Fourier components of the measured deformation of
real tires revealed that the stretched string type of model is more suitable
for the simulation of a bias ply tire, whereas the beam model is more ap
propriate for the radial ply tire. It appears, furthermore, that the use of a
stretched string model requires a tension force which is of the order of
25000 N, whereas measurements of Hinton [14] indicate that the tension
force amounts to about 2000 N. This implies that apparently the stiffness
of the carcass (shear and bending) is responsible for the rest of the effec
tive tension force.
In reference [23] Savkoor enunciates his theory for the development of a
general mathematical model for the description of the lateral behavior of
the tire. He uses results of experiments carried out with a tire which rolls
slowly at a constant slip angle. The principle of using influence (Green)
functions, describing the lateral elastic field of the tire, has also been
adopted by Pacejka [60, 71] in attempts to generate tire characteristics.
Figure 9.5.3 (taken from [12]) presents the calculated characteristics of
several types of models. The parameters in cases a, b and c are chosen in
such a way as to give a best fit to experimental data for the cornering force
at small slip angles. Curves d show the result when carcass elasticity has
been neglected and only the flexibility of the tread elements is taken into
account (Fromm [9]). When the elastic constant of Fromm's model is cho
sen in a similar way, no difference between Fromm's and Fiala's results
appear, since due to Fiala's approximations the coefficients in the ex
pressions for Fy(α) and Mz(α) are equal to those obtained directly by
Fromm.
In the calculations for figure 9.5.3, Frank applied a vertical force distri
bution qz(x), found from measurements, lying between a parabolic and an
elliptic shape. The positive aligning torque obtained at high values of a
(fig. 9.5.3) arose due to a slightly asymmetric shape of qz(x) introduced in
those calculations. The phenomenon that in practice the aligning torque
indeed varies in this way is probably due to a combination of several ef
fects. Apart from the cause just mentioned above, the rolling resistance
force acting out of the wheel plane, due to the lateral deformation, may
contribute. Another important factor causing the moment to become posi-
792 MECHANICS OF PNEUMATIC TIRES

2000

1000

10-

FIGURE 9.5.3. Comparison of calculated characteristics for the cornering force f"/vj and the
aligning torque (-MJ versus slip angle a (from Frank [12]).
Coefficient of friction p constant. Vertical pressure distribution qz slighily asymmetric The curves stand for a. stretched
beam model, b. beam model, c. approximation of Fiala, d. model of Fromm (rigid carcass).

tive is the fact that the coefficient of friction is not a constant but depends
on the sliding velocity, the latter having its highest values in the real por
tion of the contact area. This factor may also cause the slight drop in the
/"»(α) curves as has sometimes been found experimentally at high values of
α. The influence of different but symmetric shapes for the vertical force
distribution along the jc-axis has been theoretically investigated by Bor-
gmann [16]. He finds that, especially for tires exhibiting a low carcass stiff
ness, the influence of the pressure distribution is of importance and has, as
may be expected, particular effect on the aligning torque at higher values
of the slip angle. Many authors adopt the parabolic distribution for pur
pose of mathematical simplicity (Fiala [13], Freudenstein [4], Bergman
[17], Pacejka [18]).
Figure 9.5.3 shows that, when the model parameters are chosen prop
erly, the choice of the type of carcass model hardly influences the results.
For illustration, we shall present now the theory of steady state side slip
with the aid of the simple model of Fromm (cf. [9]), and the more ad
vanced model of the stretched string type with and without tread elements
ANALYSIS OF TIRE PROPERTIES 793

(for details cf. Pacejka [19]). These two examples were chosen for reasons
of their connection with theories to be presented later on concerning the
influence of a driving or braking force and nonsteady state tire behavior,
respectively.
Tire model with elastic tread elements and rigid carcass (Fromm)
The model to be treated first is shown in figure 9.5.4. The steady state
drifting tire shows a contact line which is straight and parallel to the veloc
ity vector V in the adhesion region, and curved in the sliding region where
the available side force becomes lower than the force which would be re
quired for the tips of the tread elements to follow the straight line further.
In the adhesion region the shape of the deformation is in accordance with
the general equation (13). It is easy to prove that at the leading edge the
deformation of the tread elements vanishes. Consequently the lateral de
formation in the adhesion region reads, when the drift angle α is assumed
to be relatively small so that we can write tan α ≃ α:
^i^'-^ <9-5-15)
where a denotes half the contact length.
In case of vanishing sliding, which will occur for o —» 0 or for ju —> oo,
expression (9.5.15) holds for the entire region of contact. After the in
troduction of the total lateral stiffness cP of all profile elements per unit
length of the assumedly rectangular contact area, the following ex
pressions for the cornering force Fy and the aligning torque -Mz are ob
tained:

(9.5.16)
-M, = -cf vxdx

The cornering stiffnesses for the force and the moment consequently be
come respectively:

(9.5.17)

rigid carcass
tread rubber fM

TOP-VIEW

FIGURE 9.5.4. Simple model of drifting tire (Fromm).


794 MECHANICS OF PNEUMATIC TIRES

We will consider now the case of a finite value of pi and a pressure distri
bution which gradually drops to zero at both edges. For purpose of sim
plicity we assume a parabolic distribution of the vertical force per unit
contact length as expressed by

(9.5.18)

where W represents the vertical wheel load (=-Fz). The largest possible
side force distribution consequently reads:
ef-x2
(9.5.19)

In figure 9.5.4 the maximum possible lateral deformation νmax = yr^f


has been indicated. We introduce for the sake of abbreviation the follow
ing tire model parameter

(9.5.20)

The distance from the leading edge to the point where the transition from
the adhesion to the sliding region occurs equals 2aλ and is determined by
the nondimensional quantity λ, which bears the following relation to the
slip angle α (assumed positive):
\ - 1 - 9a. (9.5.21)
From this equation the angle αsl can be calculated, at which total sliding
starts (λ = 0):
ad=\/0. (9.5.22)
The force Fy and the moment Mz can now be derived easily as a function
of α. The results read:
- X3) -
for a < <*„ (9.5.23)

for a,, < a < —

-M. - X) - 0a - 3(6a - 3(6a?

for a < aw (9.5.24)


<ff
J/,-0 for aw < a < -T-

These relationships are shown graphically in figure 9.5.5. They correspond


with curves c or din figure 9.5.3, but now for a symmetric parabolic pres
sure distribution. The pneumatic trail t, which indicates the distance be
ANALYSIS OF TIRE PROPERTIES 795

hind the contact center C where the resultant lateral force acts, becomes at
vanishing slip angles:
M,\
Via. (9.5.25)
CF.

This value is smaller than normally encountered. The introduction of an


elastic carcass will, as we shall see, improve this quantitative aspect.
Another point in whrch the simple model deviates considerably from
experimental results is the effect of a variation of the vertical wheel load
W. With the assumption that a changes quadratically with W, it can easily
be shown that Fy and M will vary proportionally with Wand W*n respec
tively. As will appear later in this respect also the introduction of an elastic
carcass improves agreement with experiment.
It should be noted that Fiala, who obtained the same expressions
(9.5.23-24) via approximation of his more complex relations derived from
a model with both beam and tread rubber, succeeded in finding suitable
values of his original model parameters (combinations of which govern
the parameter comparable with our quantity θ) so that reasonable agree
ment with experimental curves are obtained including the variation with
wheel load W. This latter relation Fy(W) even appears to show a maxi
mum, after which Fy drops with increasing W. An investigation is needed
as to whether this drop is due to the approximation of a symmetric para
bolic shape of carcass. It should be noted that with Fiala's model a drop of
/', versus W is accompanied by a region of negative lateral force distribu
tion in the forward portion of the contact line. This force distribution does
not appear to be possible for stretched string tire models. Another theory
explaining this drop, which has been observed in many experiments
(Gauss and Wolff [2]), makes use of the reduction of the lateral stiffness of
the carcass as a function of vertical deflection, due to variations in geome
try of the cross section of the tire in the contact region (cf. Rotta [20] for
theory, Smiley and Home [8] for experimental verification and Bergman
[17] for application).
Curves obtained in the experiments of Freudenstein [4] for truck tires
and Nordeen and Cortese [5] for passenger car tires do not show such a
maximum in their ranges of measurement. As an illustration, some experi
mental results are shown in figures 9.5.6-8. In these figures different ways
-Mil Fr

6 a (rad)

FIGURE 9.5.5. Cornering characteristics of the simple model with rigid carcass (Fromm).
796 MECHANICS OF PNEUMATIC TIRES

of plotting are shown, each of which have their specific advantages. The
curves of figure 9.5.6 may be more suitable for use in the analysis of auto
mobile motions. Both the functions Fy(α) and F1W) are directly obtain
able from this kind of carpet plot. For further information we refer to part
8.3 which contains an extensive collection of experimentally determined
cornering force and aligning torque characteristics (figs. 8.3.22-35). In or
der to obtain a deeper insight into the problems related to side slip, we
shall turn now to the treatment of a more advanced tire model.
Stretched string model with tread elements
The analysis to be presented now has been taken in an abridged form
from Pacejka [19]. We shall start with a more general analysis which will
also cover the possibility of antisymmetric longitudinal deformations oc
curring in the shimmy motion to be dealt with in section 9.5.3. In figure
9.5.9 a top view of the model is shown in an arbitrary position. The carcass
is represented by a number of elastically supported parallel strings under
tension, which are connected by cross cords. The points of connection can
move only laterally, and their mutual distance remains the same. When

-M

1700 N

FIGURE 9.5.6. Carpet plots of cornering characteristics for 7.60-15 tire (presumably
conventional cross-ply), p, « 2 bar.
Dry flat road surface, low >peed (from Nordeen and Cortex |S]).
ANALYSIS OF TIRE PROPERTIES 797
Normal

35

100 200 300 MO 500 600


-M,
Aligning torque (Mm)

FIGURE 9.5.7. Cough-plot of truck cross-ply tire 9.00-20 eHD, p, - 5.5. bar.
Dry road measurements. V - 10 Km/h (from Freudenstein, Ref (4]).

the strings are deformed laterally the rubber between the strings will be
sheared. Through the continuous elastic support, axial forces distributed
over the length of the band can be transmitted to the wheel plane. To this
band under tension, several rows of an infinite number of elastic blocks
are attached, representing tread elements. In contact area of length 2a and
width 2b the ends of these elements are in contact with the road surface.
The strings are assumed to be of infinite length.
The longitudinal deformation u is assumed to be proportional to the
longitudinal component of the contact pressure. The following relation
holds:

pneumatic trail t (mm)

wo »o
-M2 (Nm)
Aligning torque
FIGURE 9.5.8. Gough-plot of 6.00-13 Dunlop tire, p, - 1.4 bar, dry internal drum with inner
diameter 3.8 m, speed V — 40 km/h (from Henker [6]).
MECHANICS OF PNEUMATIC TIRES

wheel axis

string
crosscord

FIGURE 9.5.9. Top-view of tire model considered

P* = -c,,u, (9.5.26)
where p, is the force and cpx the longitudinal stiffness of the profile ele
ments, both per unit area.
The lateral deflection v is made up of the lateral deflection of the string
(the carcass) νc and the lateral deflection of the tread rubber v,:
v - vc + v,. (9.5.27)
We will consider only the case where v,, is constant along the width of the
contact area, as will occur in cases to be investigated. We assume νp to vary
proportionally with qy as expressed by:
2bpy = qy = -Cfvp (9.5.28)
where qy denotes the lateral force and cp the lateral stiffness of the tread
elements, both per unit length.
In order to obtain an expression for the deflection of the strings we must
consider the equilibrium of an element of the tire model as shown in fig
ure 9.5.10, where the longitudinal displacements u, resulting in a second-
order effect, are neglected. In the lateral direction, the equilibrium of
forces acting on the element of length dx and full tread width 2b results in
the following equation:

D-D- dx - 5,

= 0, (9.5.29)

where cc denotes the carcass stiffness ("pneumatic stiffness") per unit


length (cf. eq (9.4.6), fig. 9.4.2), S1 the longitudinal component of the total
tension force in the strings and D the shear force in the cross section of the
tread-band. The shear force is assumed to be a linear function of the shear
angle, according to the formula:

D = -52|^- (9.5.30)
ANALYSIS OF TIRE PROPERTIES 799
With the introduction of the constant S = S1 + S2 we deduce from eq
(9.5.29):

(9.5.31)

In the part of the tire not making contact with the road the contact pres
sure vanishes so that:

(9.5.32)

For that part making contact with the road we obtain, with eqs (9.5.27-28,
9.5.31)

S —y - cfvc = - c,(v - vc) for |jc| < a. (9.5.33)

We introduce the tire constants:

-& °--^- ™d<-7-MA: <"•34)


With increasing tread rubber stiffness the value of the parameter € de
creases until it vanishes when cp —» oo, which represents the case of no
tread elements. When the quantities denned in eq (9.5.34) are introduced,
one obtains the equations of equilibrium (9.5.31), (9.5.32), and (9.5.33) as

o'Sf-Vc-^c, (9-5-35)

(9.5.36)

FIGURE 9.5. 10. Equilibrium of deflected tire element.


800 MECHANICS OF PNEUMATIC TIRES

o? r - *, = -(!-*> for |*|<fl. (9.5.37)

For large values of |x| the deflection v, tends to zero. Therefore the solu
tion to eq (9.5.36) reads:
vc = Cf" for x> a
(9.5.38)

At boundaries x — ± a we obtain:
i- av e
vc — —a lim —— for x = a
(9.5.39)
,. dv
vc = a lim — for x — -a.
xt-. ox

Since for x = ±a the deflection v,. and its derivative dvjdx vary continu
ously with x, the latter due to the fact that no finite concentrated forces
can act on the strings, with finite deflection νp and finite stiffness cp (cf. also
eqs (9.5.28) and (9.5.31)), in the expressions (9.5.39) the limit signs may be
omitted, after which they can be used as boundary conditions for the solu
tion of equation (9.5.37). For the determination of the integration con
stants occurring in the solutions of the first-order partial differential equa
tions (9.5.9), the additional conditions are needed that the deflections ν
and u vary continuously at the leading edge, where x = a. That this conti
nuity does take place can be proved in the following way.
For the real tire, where /i is finite and the vertical pressure gradually
tends to zero at the leading and trailing edges of the contact area, it will be
obvious that the tread elements show no deflection just after entering the
contact area or just before leaving this area. Consequently the deflections
vary continuously in the neighborhood of both edges in this case.
For the extreme case where finite shear stresses are available at the lead
ing and trailing edge of the contact area (ju -» oo) a finite deflection may
occur at these edges. It can be shown, however, that if we consider van
ishing regions of sliding at both edges, at the leading edge sliding veloci
ties would occur which produce friction forces directed opposite to the ex
ternal forces required for maintaining the discontinuity, whereas at the
trailing edge this will not occur. Similar findings are obtained for the
model without tread rubber when kinks are assumed in the string at both
edges. The conclusion must be that only at the trailing edge may a finite
change in deflection of the tread rubber, or a kink in the string without
tread rubber, occur in the extreme case as /i —» oo (cf. [19] for detailed dis
cussion).
The forces and the moment acting on the tire may be computed by in
tegration over the contact area A. The forces in the longitudinal and lat
eral directions become respectively:

- p,4A. (9.5.40*)
ANALYSIS OF TIRE PROPERTIES 801

The moment about the vertical axis reads:

«. - / (P*y - P,x)dA. (9.5.40*)


JA

In case of purely lateral slip the tire model does not show longitudinal de
formations, so that px = Fx •• 0. The lateral force and the moment then,
become:
"" (9.5.41)

The general analysis derived above will be applied here for the case of
steady state rectilinear side slip. As before, the vertical force distribution is
assumed parabolic along the x-axis. We obtain for the lateral force distri
bution in the region of sliding:
(9.5.42)
where

' (9-5-43)
When we reduce the slip angle from a large value, where total sliding oc
curs, we find a point of first adhesion which is situated somewhat behind
the leading edge when the model parameter e is smaller than a certain crit
ical value depending on σ, i.e., when the lateral stiffness of the tread rub
ber exceeds a certain value. This means that when reducing the slip angle
further, two regions of sliding occur: a small region in front of and a much
larger one behind the region of adhesion. The critical value of e above
which adhesion occurs immediately at the leading edge is given by the fol
lowing formula
1- tanh
(9.5.44)

We shall discuss here only the results for a model with relatively low tread
rubber stiffness exhibiting only one sliding region, and for a model with
out tread rubber elements (cp -» oo) showing two sliding regions. Figure
9.5.1 1 shows the deflected tire model of the former kind. Differential equa
tions (9.5.13) and (9.5.37) are applicable for the adhesion region (a2 < x <
a) and (9.5.35) and (9.5.43) for the sliding region (-a < x < a2). There are
five constants of integration and one unknown a2. We therefore need six
boundary conditions in order to find these.
These conditions follow from (9.5.39) and the discussions thereafter.
The slip angle α and the shape of the deformation are calculated as a func
tion of the distance a2 for the special case e = 1/7.5 and σ = 3.74 a. Figure
9.5.12 shows the shape for a number of cases. The obliquely shaded area
indicates the sliding regions, which grow with increasing slip angle until
802 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5. 1 1 . Deflected tire model provided with tread elements, showing an adhesion and a
sliding region.

the whole contact line slides. Larger slip angles will not alter the shape
anymore.
The relaxation length σ* has been defined as the distance between the
leading edge of the contact area and the point of intersection of the elon
gation of the straight portion of the contact line with the x-axis. The val
ues of σ and e mentioned above were chosen in such a way that σ* tends to
the value 3a for a —» 0. Note that the relaxation length decreases from the
value 3a to a value somewhat below 2a when total sliding starts. This
property has been confirmed experimentally by Metcalf [55].
Once the deflection of the tire model is known, the force and moment
can be calculated by the use of eqs (9.5.14) and (9.5.28). Integration over
the contact length as indicated by eq (9.5.41) yields expressions for /',. and
Mz in terms of a2. Figure 9.5.13 shows the calculated tire characteristics.
As with Fromm's model, the slope of both curves becomes zero at the slip
angle a,, where total sliding starts. We shall see that this is not the case
when the tread rubber is removed from the string (cp —» oo, e = 0, σ* = σ,
νp = 0, ν = νc).

O/i

FIGURE 9.5.12. The tire model drifting at various slip angles a (a • 3.74<j, e - 1/7.5).
The shaded area indicates the regions where sliding occurs.
ANALYSIS OF TIRE PROPERTIES 803

FIGURE 9.5. 13. Tire cornering characteristic for the tire model with tread rubber (a - 3.47a,
« - 1/7.5).

String model without tread rubber elements


Figure 9.5.14 shows this simple string model in a deflected situation.
Two regions of sliding are expected to occur. In the region of adhesion
(a2 < x1 < a1) the stationary version of eq (9.5.13) holds. The sliding re
gions are governed by eqs (9.5.35) and (9.5.42). In the straight portion of
the contact line where adhesion occurs, the following inequality holds ac
cording to eqs (9.5.35), (9.5.13), (9.5.42), and (9.5.43):

?<»•=• J 1 - (9.5.45)

with the parameter

(9.5.45a)

This means that the straight portion of the contact line lies inside the para
bola as indicated in figure 9.5.14. The points of inflection of the contact
line are located on this parabola when Vsu < 0 in these points. Near the
edges x = ±a the available lateral force tends to zero. Since a finite deflec
tion ν > 0 is present in these places, the curvature of the string will be con
cave (d2ν/dx2 > 0) according to (9.5.35). In the rear portion of the contact
line we have dv/dx > 0, so that according to (9.5.14) Vv < 0. Until the
contact line intersects the parabola, the shape remains concave. Inside the
parabola, however, the curve becomes convex. The boundary of the adhe
sion region is at x = a2. The point of transition to the front sliding region is
denoted by x = a1. When this latter point lies inside the parabola, accord
ing to (9.5.35), the curve must be convex just in front of that point when
the sliding velocity is negative. This shape, however, would lead to an in
crease in slope, so that according to (9.5.14) the sliding direction becomes
positive in that case, which is in contradiction with the assumption. In the
same way a concave shape can be shown to be impossible. The conclusion
must be that this point of transition must lie on the parabola. The curve in
the front sliding region can only be concave, so that Vn < 0. Its curvature
804 MECHANICS OF PNEUMATIC TIRES

_a ^ a_

FIGURE 9.5. 14. Deflected tire model without tread rubber showing two sliding regions.

tends to zero when the parabola is approached. This forms one of the
seven boundary conditions necessary for the calculation of a1 and a2 and
the five constants of integration in the solutions for the three differential
equations. By means of iteration the contact line has been computed for a
number of values of the slip angle α. Figure 9.5. 15 shows the results. The
variation of the relaxation length with slip angle will be extremely small
for the model considered.
The cornering force and the aligning torque are found by integrating
the external lateral forces along the contact line. The results of the compu
tations are shown in figure 9.5.16. In contrast to the characteristics of the
model with tread elements shown in figure 9.5.13, the curves show a dis
continuity at the slip angle where total sliding starts.
Behavior at vanishing sliding
It is of interest to know the behavior of both models with and without
tread rubber in case of vanishing sliding, i.e., for coefficients of friction
tending to infinity or for slip angles tending to zero (see fig. 9.5.17). The
model without tread rubber and with vanishing sliding was originally

lit

8, iv 002 7
Irani sliding ragfeni
•dhnton n>gion»
r««r sliding regions

FIGURE 9.5.15. The tire model drifting at various slip angles a(o - 3o, « - 0).
The shaded areas indicate the regions of sliding.
ANALYSIS OF TIRE PROPERTIES 80S

FIGURE 9.5.16. Tire cornering characteristics for tire model without tread elements
(a = 3a, t - 0).

treated by von Schlippe [9, 25, 26] and Temple [9]. Their theory is of par
ticular importance for the study of nonsteady state motions as treated in
section 9.5.3.
Omitting the detailed calculations, we obtain for the lateral deflection at
the forward contact point:
v, = <j*a (9.5.46)
where
"•c-2} -4a
ff* = (9.5.47)
1+e 1 -e
1 -c 1 +e
which is the expression for the relaxation length at zero slip angle as pre
viously denned (see figure 9.5. 17a). The relaxation length equals a when
the tread stiffness tends to infinity (see fig. 9.5.17b).
The cornering force and aligning torque become in this case:

Fy - CFaa and - M, = CMaa (9.5.48)


806 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5. 1 7. Drifting tire models at vanishing sliding; a. with tread rubber, b. without tread
rubber.

with the stiffnesses CFα and CMα having the following values:
CF, - 2cc(l - e2) [(a* + a)a - lAoa* {(1 + &"•• + (1 - e)<rw<>< - 2}

+ Wo^l - e2) (<?"°< + e-*°"< - 2)],


(9.5.49)
CHa - 2c,(l - e2) ['/so3 - '/40{a«(l + e) - a(l - e2)}
+ ac(l- e""<)} - l/4o{a*(l - €) - a(l - e2)} {a

in which σc and e are denned by (9.5.34). When the tread elements are
omitted (αc = e = 0), these equations reduce to:
Cfa = 2cc(a + a)2 and CMa = 2c^j {o(a + a) + Via*} . (9.5.50)
If the contact patch of a standing tire is moved sideways a lateral force
will arise, which for small deformations is proportional to this lateral shift
according to the relation

where C\ denotes the lateral stiffness of the tire and v,, represents the lat
eral deformation of the contact patch relative to the wheel rim. In case of
side slip, /', assumes the same value when the center point of the contact
line shows a lateral deflection equal to v,,. When taking into account that
ν0 = (a + σ*)α the following relation appears to hold for the relaxation
length of a tire

— a. (9.5.51)
ANALYSIS OF TIRE PROPERTIES 807

This expression may be used for the experimental determination of the re


laxation length σ* of a real tire.
For a number of values of a and cp/cc the resulting relaxation length σ*,
pneumatic trail t0(=CMJCFα), and cornering stiffness CFα are calculated
and listed in table 9.5.1. The original value of the lateral carcass stiffness
per unit length cc has been designated with cco. It is remarkable that the
inclusion of tread rubber of a relatively high stiffness (cp/cc = 55) does re
duce σ*, t0 and CF so much. The model with tread rubber may give results
close to those obtained experimentally (cf. Fonda and Radt [7, figs. 25,
34]).
In figure 9.5.18 for the same combinations of parameters (except σ = 3a)
the variation of σ*, CFα and CMα with vertical wheel load Wis shown when
a parabolic variation of the contact length 2a with W is assumed. The
original values are designated with «„ and W0. It is shown that relative to
the behavior of Fromm's model (cc —» oo), the variation of the cornering
stiffness with wheel load has improved somewhat due to the introduction
of carcass elasticity. The case without tread rubber (c.—» oo), in turn, dif
fers qualitatively from reality since the cornering stiffness Cr,, remains fi
nite for the vertical load W equal to zero. The drop sometimes observed at
large values of W cannot be simulated with this model unless a cc is varied.
It is then necessary to introduce two different functions of average lateral
foundation stiffness ccF( W) and ccM(W), one for the force and the other for
the moment. This is plausible since for the force the lateral deformation of
the center portion of the contact area is of primary importance, while for
the moment the portions outside the contact area play the greatest role,
and these show less radial deflection. As has been shown in section 8.3.1
the stiffness cc decreases with increasing radial deflection. The variation of
the relaxation length σ* as indicated in figure 9.5.18 is of importance for
the analysis of section 9.5.2, where the influence of a time-varying load is
treated.
Camber and Turning
For antisymmetric steady state motion of automobiles, one must con
sider not only side slip but also the influence of two other effects, which in
most cases are of much less importance than side slip. These two variables
which complete the description of the antisymmetric motion are first the
wheel camber, or tilt angle y between the wheel plane and the normal to
the road (cf. fig. 9.5.19), and second the spin or degree of turning. In the
steady state case the spin equals the curvature 1/R of a circular path with
radius R. In idealized form the mechanisms of camber and turning can be
considered to be similar as has been enunciated in the following.
As indicated in figure 9.5.20, the wheel is considered to move tan-
gentially to a circular horizontal path with radius R while the wheel plane
has a constant camber angle γ. The wheel is assumed to be part of an
imaginary ball. When lifted from the ground, the intersection of wheel
plane and ball forms the peripheral line of the tire. When loaded verti
cally, the ball and consequently the peripheral line are assumed to show
no horizontal deformations, which in reality will be the approximate case
for a homogeneous ball showing a relatively small contact area (fig.
9.5.20).
We apply the theory of rolling and slipping and consider equations
(9.5.4). We assume:
808 MECHANICS OF PNEUMATIC TIRES

S 2
o. —

s
if

<*
JJ
to
I1
HI

ead a s
" 3
w O
!2
•c
1 J3 £

e
H g,g,

II

a a
s s
809
ANAUVS1S OF TIRE PROPERTIES

(9.5.52)

*L
dx
810 MECHANICS OF PNEUMATIC TIRES

whtel
axle

wheel plan*

imaginary ball

FIGURE 9.5.19. Camber and turning without side dip.

For the combination of turning and camber u, reads:


V
u, - «„ + «„ - — - ft sin Y- (9.5.53)

The total spin <f> consequently becomes (cf. eq (9.5.8):


u. 1 1 (9.5.54)

with r. denoting the effective rolling radius. In case that complete adhesion
occurs in the contact area eqs (9:5.52) become:
du_ *L (9.5.55)
dx dx
Integration yields the following expressions for the horizontal deforma
tions in the contact area:
u — —<j>yx + constant
(9.5.56)
v «• Vitjtx2 + constant.

contact
area of
imaginary

crtlcal
projection
Of
horizontally
undcformed
peripheral lint
TOP.VIEW

FIGURE 9.5.20. Top-view of peripheral line of nonrolling tire considered as a part of


imaginary ball pressed on flat surface.
ANALYSIS OF TIRE PROPERTIES 811

The constants of integration follow from boundary conditions which de


pend on the tire model employed. As an example, consider a simple model
corresponding to that of Fromm with horizontal deformations through
elastic tread elements only. The contact area is assumed to be rectangular,
of length 2a and width 2b, and filled with an infinite number of tread ele
ments. In figure 9.5.21 three rows of tread elements are shown in the de
formed situation.
For this model the following boundary conditions apply:
x = a: v = u ™ 0.
With the use of (9.5.56) the formulae for the deformations read:
u-y(a-.
(9.5.57)
v = —l
After the introduction of cpx and cpy denoting the stiffness of the tread rub
ber per unit area in s and y direction respectively, we obtain for the lateral
force and the moment about the vertical axis, by integration over the con
tact area:

(9.5.58)
A/ =-
or in terms of camber and path curvature (γ small):
1 1
-c*
(9.5.59)
i--c^-Hi. 1-
«r — 1-
^T — -

horizontally
contact undef or med
area peripheral
I ine

path

FIGURE 9.5.21. Top-view of cambered tire model rolling in a curve with radius R.
812 MECHANICS OF PNEUMATIC TIRES

These expressions indicate that camber and path curvature have opposite
effects when their signs are equal, according to the sign convention. In
case of pure turning, the force on the tire is directed away from the path
center and the moment acts opposite to the sense of turning. Consequently
both the force and the moment try to reduce the curvature 1/R. In case of
pure camber, the force on the wheel is directed towards the point of inter
section of the wheel axis and the road plane, while the moment tries to
turn the rolling wheel towards this point of intersection. No resulting force
or torque is expected to occur when -y = re/R, which is approximately the
case when the point of intersection and the path center are the same. A
number of authors explain the presence of a moment by considering two
wheels rigidly connected to each other on one axle. In a curve the wheel
centers travel different distances in a given time interval and when cam
bered, these distances are equal but the effective rolling radii are different.
In both situations antisymmetric longitudinal creep must occur which pro
duces the moment.
Up to now we have dealt with the relatively simple case of complete ad
hesion. When sliding is allowed by introducing a limited value of the coef
ficient of friction ju, the calculations become quite complicated. If, as be
fore, a parabolic pressure distribution is assumed, it is obvious from eq
(9.5.57), that for an infinitely thin tread (b —> 0) no sliding will occur up to
a certain critical value of spin <£,,, where the adhesion limit is reached
throughout the contact length. Up to this point Fy varies linearly with <>.
According to eq (9.5.54), spin due to camber theoretically cannot exceed
the value 1/re, so that this discussion will be limited to the case of turning
when higher values of spin are considered. Beyond the critical value <£„ the
situation becomes quite complex, as has been pointed out by Freudenstein
[4]. In the front half of the contact line sliding will occur, whereas behind
the contact center adhesion takes place up to a point B (cf. fig. 9.5.22)
where the deformation ν is opposite in sign and reaches a maximum, after
which sliding occurs again. With increasing spin <J>(= l/R) this latter slid
ing zone grows while at the same time the side force — F, decreases and
the torque -Mz, which arises for </> > <J>W, increases until the situation has
been reached where R and Fy vanish and -Mz becomes maximum.
When a finite width 2b is introduced, complete adhesion will only occur
for vanishing values of spin. We expect that sliding will start at the left
and right rear corners of the contact area, since in these points the avail
able horizontal contact forces reduce to zero and the longitudinal defor
mations u would become maximum for μ —» oo. The zones of sliding grow
with increasing spin and will thereby cause a less than proportional varia
tion of -Fy and -Mz with <?>. The maximum lateral force will be lower
than in case of line contact. A more or less exact theoretical treatment of
the spin behavior of pneumatic tires with sliding taken into account has
not been found in the literature. This behavior has been treated in great
detail by Kalker [21] for the problem of rolling contact between elastic
balls. Freudenstein [4] presents a rough theory for the turning behavior up
to an approximate value of <>,,. He treats the longitudinal and lateral de
formations as being uncoupled and introduces an effective coefficient of
friction Vi n >/2. In addition, Freudenstein gives an approximate value of
the maximum torque occuring at R = 0. For a parabolic pressure distribu
tion in x-directipn and a uniform distribution in ^direction, he obtains for
contact areas with dimensions in the range b < a < 2b:
\MZ _J « tyi Wft a + b). (9.5.60)
ANALYSIS OF TIRE PROPERTIES 813

circle parabola
(adhesion) (sliding )

FIGURE 9.5.22. Simple tire model in sharp curve.

Figure 9.5.23 (from [4]) gives the turning characteristics of a truck tire. A
comparison with experimentally obtained cornering characteristics (cf. fig.
9.5.7) indicates that the values of both the lateral force and the moment
reached at one degree slip angle roughly equal the force and moment ob
tained with the same tire when rolling in a curve with R = 3.5 m. (accord
ing to experiments with bias ply truck tires). Reference [4] indicates, fur
thermore, that the cornering stiffnesses CFα and CMα are 20-30 percent
higher for the radial ply (belted) tire. Analogously, the radial ply tire is ex
pected to show greater resistance against turning than the diagonal ply
tire.
From the discussion above it follows that the effect of spin due to a sta
tionary turning may only be of importance with slow city driving. In theo
hormii whcil load W

•r, nooo4—fc*=?>

1000

4000

Nil-mil wheel load W


<Nm)

Radius of curvature R (m)


FIGURE 9.5.23. Measured turning characteristics for truck tire.
Literal force Fy and moment M, about vertical axis u a function of radius of curvature R. Tire crow-ply 9.00-20 eHD: p,
• 3.3 bar. V - 1-3 km/h on dry road (from Freudenstein [4]).
814 MECHANICS OF PNEUMATIC TIRES

retical investigations of automobile motions normally met in the literature,


this effect has been neglected. Camber, however, which is the other vari
able associated with the spin effect, will be of particular importance for
two-wheeled vehicles. The theoretical relationship between turning and
camber has been used by many authors (Fiala [13], Freudenstein [4],
Bergman [17] and Maier [9]). In fact, however, this relationship is only ex
act in the case of a rolling ball. A cambered tire will behave somewhat dif
ferently since there exists a combination of radial and lateral stiffnesses of
the tire which in general are not the same. We expect therefore that, unlike
the ball, a cambered wheel loaded vertically through a purely vertical mo
tion of the wheel axle (in which position the vertical projection of the pe
ripheral line remains approximately unchanged), will show an initial hori
zontal lateral force before the wheel starts to roll. The camber force of a
steady rolling tire may be obtained by superposition of the initial lateral
force produced by the vertical displacement, and the change of this force
during rolling due to the spinning action as treated in the theory above. In
the final steady state stage, the cambered tire will show a straight contact
line parallel to the wheel plane if sliding can be neglected. The complete
process is very complicated and a fundamental treatment of the behavior
of a rolling cambered wheel is needed.
In addition to the lateral (horizontal) force due to axial (i.e., normal to
the wheel plane) deformation, we should take into account in such a the
ory the horizontal lateral force which arises when a standing tire is tilted
about the line of intersection of wheel and road plane. From the difference
in curvature of the tire side walls in the tire cross section, Rotta [20] calcu
lates the side force which is necessary to balance the internal air pressure.
Rotta states that this force is the main contribution to the camber thrust.
Once the tire rolls, axial deformation may occur due to, amongst other
things, the torsional stiffness of the carcass and the finite width of the
tread. These factors are responsible for the so-called overturning couple
Mx about the .v-axis (intersection of wheel and road plane).
For a tire model with line contact (for instance a single row of tread ele
ments or a single elastically supported stretched string) there is no need for
the peripheral line to move out of the wheel plane when it is tilted. The
camber force and moments then simply read:

(9.5.61)

These relationships appear to hold approximately for diagonal ply tires.


Experiments indicate that radial ply tires produce less camber thrust than
bias ply tires. This is in complete disagreement with the results deduced
for turning, which reinforces the suspicion that the mechanism of camber
and turning are not completely equivalent. As Freudenstein did not give
the camber characteristics of the truck tires on which the turning behavior
was measured, we are not able to compare the responses to camber and
turning. According to Hadekel, for aircraft tires the lateral force due to
turning is about four times higher than the camber force at equal values of
spin. Evaluation of Freudenstein's truck tire data reveals that for bias ply
tires (assuming that relation (9.5.61) holds for these tires) a lower ratio of
about 1 is expected; for radial ply tires, however, Hadekel's value may be
of the right order of magnitude. In figure 9.5.24 some experimental cam
ANALYSIS OF TIRE PROPERTIES 815

her characteristics are shown. We furthermore refer to figures 9.3.37 and


9.3.38.
Influence of Braking and Traction Forces
According to experiment large longitudinal forces have considerable in
fluence on tire cornering force and aligning torque. Theory on this subject
is scarce. A significant paper is that of Bergman [17] which, however, is re
stricted to the influence of traction. Bergman had to simplify the system
considerably, due to the exceedingly complex mechanism of the real tire.
He employs the concept of interaction effect and effective lateral coeffi
cient of friction. Bergman states that traction reduces the lateral stiffness
of the standing tire, resulting in a reduction in cornering force. Similarly,
one would expect that braking also would soften the tire in the lateral di
rection. However, experiments indicate that moderate braking increases
the cornering force slightly. This contradicts Bergman's concept. The in
creasingly sharp reduction of the cornering force at high values of traction
has been explained by Bergman by the introduction of the effective coeffi
cient of friction in the lateral direction μy = V/*2 ~~ M« where n, represents
the effective tractive coefficient /•", / W. It is obvious that this is a crude ap
proximation to reality, although the results appear to fit some observa
tions. However, for the description of the influence of the braking force,

ilOON
MOON
i.rmalwh.tl (••*, W
(DON

in ON

•MM
Normal whi«l lMd,W
.4*00 N

2700N

FIGURE 9.5.24. Measured camber characteristics.


Lateral force F, and moment M, u a function of camber angle y and wheel load W. Same tire ai in figure 9.5.6 (a - 0)
(from Nordeen and Cortex (51).
816 MECHANICS OF PNEUMATIC TIRES

and especially for the influence of longitudinal forces on the aligning


torque, Bergman's theory is inadequate.
A theory which is probably more widely applicable will be presented
next. It is a further development of a first attempt of the author reported in
[18]. It describes the behavior of a relatively simple model, which later on
will be extended to a more advanced model.
Model without carcass elasticity
The simple model is identical to Fromm's model, for which the steady
state side slip behavior has been treated in section 9.5.1. The carcass is
considered to be rigid, and the tread elements provide the necessary elastic
properties of the tire. For purposes of mathematical simplicity it is as
sumed that the horizontal stiffnesses of the tread elements, of which sev
eral rows may be present, are equal in the lateral and longitudinal direc
tion. We introduce:
cp = 2bc,» = 2bcry (9.5.62)
with cf, and <;,„ denoting stiffnesses per unit area, and cp the horizontal
stiffness per unit of length.
In figure 9.5.25 the deformations of the tread elements are shown for
both the cases of traction (Fx > 0) and braking (Fx < 0). The longitudinal
creep velocity Vex obviously becomes:
V., = Fcos a - flre (9.5.63)
The longitudinal slip ratio has been defined as

The deformations due to longitudinal slip and side slip become, in case of
full adhesion (fi —» oo):
u = (a — X)K
(9.5.65)

where for simplicity α is assumed to be small. We introduce the angle δ,


indicating the angular deviation of the total horizontal force from the lat
eral direction (cf. fig. 9.5.25). The following relations hold:
sin S = FJK, cos 8 = FJK. (9.5.66)
The relation between K and a reads:

"•"-(T^T (9'5-67)
Analogous to eq (9.5.21), we solve with (9.5.20) for the point of initial slid
ing, but now with the presence of longitudinal slip (α > 0):
% . W-ic)a . „ ic . Oa (9.5.68)
cos S sin o a sin 0 + cos 8
ANALYSIS OF TIRE PROPERTIES 817

For a given slip angle α, λ will be greater at positive δ (traction) than at


negative δ (braking) for the same absolute values of δ.
When for a combination of Fx and Fy total sliding begins (λ = 0), the
slip angle and the longitudinal slip ratio are:
cos 3
jsinS. (9.5.69)
0 - sin S'

Integration over the contact length yields the total horizontal force:
K = nW(l-\3) (9.5.70)
and for the aligning torque:
M, = - iiWa\\\ - A) cosfi. (9.5.71)
From these formulae, for given α and κ, the tire forces and moment can be
calculated with the use of eqs (9.5.66-68). In figure 9.5.26 a number of
curves are shown for the parameter value θ = 5. For both the moment Mz
and the force Fy plotted against Fx a slight asymmetry appears to occur.

' inclined
parabola

FIGURE 9.S.2S. Deformations of the simple model due to side slip (a) and longitudinal
slip (K).
818 MECHANICS OF PNEUMATIC TIRES

In the (Fr , Fx) diagram curves for constant longitudinal slip values κ are
also shown. From eqs (9.5.66, 68, 70) we obtain the following formula:

For K < fiW, one finds that 0 < fte/sin δ < 1 and formula (9.5.72) in
dicates that for a given κ the absolute value of the longitudinal force |Fx|
decreases with increasing cos δ, and consequently with increasing slip
angle α.
These theoretical observations correspond to experiments carried out
by, among others, Henker [6] on a rotating internal drum (dry, cf. fig.
9.5.27) and by Holmes and Stone [27] on wet pavement. The phenomenon
that the curves on wet surfaces tend to end inwards may be due to the fact
that at higher sliding velocities the coefficient of friction decreases (cf. fig.
8.3.49).
However, the variation of the aligning torque Mz with the longitudinal
force Fx does deviate markedly from experimental results presented by
Nordeen and Cortese [5] (cf. fig. 9.5.28). As shown, the force Fy varies as
expected but the moment Mz does not agree with the theory. It appears the
Mz changes its sign during heavy braking. This phenomenon cannot be cx-

FIGURE 9.5.26. Variation of cornering force Fy and aligning torque M, as a function of


driving or braking force Fxfor the tire model shown in figure 9.5.25.
ANALYSIS OF TIRE PROPERTIES 819

F«(N) W IOM 2000 W -FX(N)

TRACTION BRAKING

FIGURE 9.5.27. Measured sideforce (F^-longitudinalforce (FJ relationshipfor constant slip


angle a (also shown for constant longitudinal slip value K).
Same tire and conditions as in figure 9.5.8. (W - 2500 N) (from tinker |6])

plained with the simple model employed so far. Before we adopt a more
advanced model, the cornering stiffnesses of the simple model will be de
rived as a function of Fx.
After elimination of X from eqs (9.5.68) and (9.5.70) we obtain, for Fy -»•

(9.5.73)

•Ml .Fy
Mm)1 IN)

FIGURE 9.5.28. Measured variation of both Fy and M, with F,for constant slip angle a.
Same tire at in figure 9.S.6 I W - 4500 N, -, - 0). (from Nordeen and Coneae |5|).
820 MECHANICS OF PNEUMATIC TIRES

which yields:

CF, = -F, + ?£1 (9.5.74)

which reduces, for /i —» oo, to


CFa^2c^-Ff. (9.5.75)
The pneumatic trail t = - Mz/Fy becomes (using eqs (9.5.66, 68, 70)):

(9.5.76,
1 - A3

or when a -» 0:

,9.5.77)

from which is obtained:


€„„ = /„ • Cfo = a8CF. IF'1 (9.5.78)
Wa • *\

and for fi -» oo;


(9.5.79)
These results are not of direct importance since they deviate too much
from experiment. They are, however, needed for the development of the
theory of the more advanced model with carcass elasticity.
Model with carcass elasticity
The simple model will now be extended by the introduction of lateral,
longitudinal and torsional elasticity of the carcass. For simplicity we shall
assume the carcass equatorial line to be straight in the contact region (see
fig. 9.5.29). For the sake of distinction the quantities referring to the
simple model of figure 9.5.25, which can be recognized as a part of the
more advanced model, are indicated by upper bars. The quantities f\, M,,
t and ( ', „ have been determined as a function of F, and d in the preceding
portion of this section.
The angles α and ft (fig. 9.5.29) are assumed to be small. With the car
cass stiffnesses Ccx, Ccy and Ccβ we obtain for the displacements of the
straight carcass section:

v0 (9.5.80)
ANALYSIS OF TIRE PROPERTIES 821

-M

FIGURE 9.5.29. Deformations of the more advanced tire model showing carcass and tread
rubber elasticity.

Furthermore the following relations hold:

(9.5.81)

a-d
and in case of complete sliding:

(9.5.82)

M, = - F,Fy J
C,
With the above equations, Fx, h\. Mz, and α can be calculated without dif
ficulty for a given combination of K and δ. By means of interpolation,
curves for constant a have been obtained, which are shown in figure
9.5.30. The calculations have been carried out using the values:
0 - 5, -3, Cff/n Wa = 7. (9.5.83)
For investigation of the influence of longitudinal elasticity, some finite val
ues of C, , are considered as well.
The Fy vs. α and Mz vs. α curves obtained for Fx = 0, figure 9.5.30c, d,
appear to be qualitatively equal to those shown in figure 9.5.5 for the
simple model. As expected, the slopes at α = 0 are less steep but the slip
angle at which total sliding starts remains the same. Therefore, the curves
in the /•', vs. Fx diagram (a) become somewhat flatter relative to the curves
of figure 9.5.26. Owing to the lateral deformation v,, the longitudinal force
Fx produces a moment about the contact center C which opposes the
aligning torque -Mz in case of braking, and thereby causes a change in
sign of the aligning torque (diagram b) as has also been observed experi
mentally (cf. fig. 9.5.28). In the same figure 9.5.30b, the influence of equal
stiffnesses C« = Ccy has been indicated. In this case the torques produced
822 MECHANICS OF PNEUMATIC TIRES
by Fx and /•; about C balance each other so that no change in sign occurs.
The influence of carcass elasticity may be smaller than predicted here be
cause of the expected property that the displacements of the lines of action
of f-\ and F are smaller than the deflection v,, and u,, respectively. Smiley
and Home [8] give some information on the shift of the normal force rela
tive to the horizontal deformations of a standing tire. They give figures of
80 percent and 25 percent respectively for aircraft tires.
The cornering stiffnesses valid for the more advanced model are derived
as follows. With the relations (81, 82) we obtain for a -» 0:

a a+p l +i
-A/, -lit, + F.V. - *>„
CM« ' (9.5.84)
a

1 Ccy Cc.

in which CFα and to are given by eqs (9.5.74) and (9.5.77). In figure 9.5.31a)
the functions (9.5.84) are plotted for the values given in eq (9.5.83), and in
addition for two finite values of Ccx. It has been found that through the in
troduction of carcass elasticity considerable qualitative changes in charac
teristics can arise.
When braking, the cornering stiffness CFα initially increases slightly, it
then passes a maximum after which a sharp drop occurs, and finally com
plete sliding takes place. From this point C,,, decreases with increasing
longitudinal slip velocity up to the point of a locked wheel. In case of trac
tion C,-,, decreases continuously and drops to zero in the range of complete
sliding, when the slip velocity - Vcx tends to infinity (wheel spin-up).
As indicated, the variation of CMα depends on the ratio of longitudinal
stiffness. For £?„ —» oo, the aligning torque stiffness CMα decreases with in
creasing braking force, becomes negative, passes a minimum and finally
shows a small negative value when the wheel is locked. In case of traction
CMα first shows a slight rise after which CMα tends to zero for Fx → μW.
The influence of the longitudinal stiffness is of great importance as
shown in the figure. In case of equal stiffnesses Ccx = C,:>, the variation of
CMα (and the moment Mz, cf. fig. 9.5.30b) becomes more symmetrical,
which corresponds to the experimental curves for radial ply tires. Bias ply
tires, however, behave according to the theoretical curves for relatively
large longitudinal carcass stiffness. The much lower longitudinal stiffness
of radial tires might indeed be responsible for the great qualitative differ
ences observed experimentally between diagonal and radial ply tires [28].
Limitations of theory
The theory just developed is limited due to .simplifications in the tire
model. Probably the most questionable simplifications are the equal stiff
nesses of the tread rubber elements in the lateral and longitudinal direc
tions, and the assumption that the carcass remains straight in the contact
zone. Furthermore, the theory assumes a constant coefficient of friction.
As has been pointed out previously, the force and moment characteristics
will change in shape due to the drop of the coefficient of friction with slid
ing velocity (cf. e.g. fig. 9.3.49). The limitations no longer appear to be es-
ANALYSIS OF TIRE PROPERTIES 823

FIGURE 9.5.30. Tireforce and moment characteristics as obtainedfor the more advanced tire
model offigure 9.5.29.

sential when the hybrid computer simulation technique, to be discussed


further below, is used. Although a number of aspects may be clarified with
the aid of the model, it is certain that important factors are not yet taken
into account. One of these will be mentioned briefly. It is the influence of
the change in tension force or effective tension force S in the carcass tread
band due to braking or driving forces.
Influence of change in tension force, front and rear
In order to investigate this influence, an alternative tire model must be
adopted. Obviously, the simple stretched string model is most suitable.
The slopes of the CFα (Fx) and CMα (Fx) curves at Fx = 0 will be calculated
while we restrict ourselves to the case of vanishing sliding (μ —* oo).
The slopes to be calculated, consequently, are purely due to the elastic
properties of the tire and may be compared with the slopes obtained with
the simple model for fi -> oo (cf. eqs (9.5.75) and (9.5.79)). In figure
9.5.31/7 the stretched string model is shown subjected to lateral and longi
tudinal forces and the aligning torque. Like in the theory of steady state
824 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5.3 la Cornering force and moment stiffness (at a — 0) as a Junction of


longitudinal force.

side slip (eq 9.5.31), the effective longitudinal tension force S is in


troduced, which, due to the longitudinal force Fx will differ now at the
trailing and leading edges. We assume the following relations to hold:
5, - S - ViFz,
(9.5.85)
S2 = S + ViF,.
According to the definition of the relaxation length when tread elements
are absent (σ, eq (9.5.34)) we obtain the following formulae for the relaxa
tion length front and rear respectively:
a? - (S - UF,)/cc , al-(S+ HF,)/cf . (9.5.86)
Analogous to eq (9.5.50) the cornering stiffnesses read:
CFa = C. {(a, + 02 + 2a)(a, + a) + a(a2 - a,)}
a) + ol(al + a) + % a2 + (a2 - a,)(o, + a)} , (9.5.87)

FIGURE 9.5.3 lb. Stretched string model deformed due to side slip and traction (no sliding.
ANALYSIS OF TIRE PROPERTIES 825

which with the aid of (9.5.68) and with the assumption F,*KS reduce to:

(9.5.88,
Ha - CMa

where index o refers to the situation without external longitudinal forces.


The slope at Fx = 0 becomes:

It may be noted that the slopes of (.",., and CMα differ in sign. The change
in tension force due to traction or braking, consequently, intensifies the ef
fect of lateral carcass elasticity.
Hybrid Simulation
A technique will now be discussed which opens the possibility of inves
tigating and simulating tire models which exceed existing models by an or
der of magnitude in complexity. The original text [60] provides more de
tails, especially on the more advanced model with carcass flexibility and
on the circuitries employed. Frank probably was the first to recognize the
potentials of following the motion of a material point of the tire [11]. Will-
umeit has followed a certain tread element during its motion through the
contact patch [70]. Essentially, this method has been employed in the cal
culation of the response of two types of tire model. The hybrid computer
appears to be the appropriate device to solve the problem.
The 'brush type' tire model to be used consists of a rigid carcass pro
vided with a row of elastic tread elements which contact the road surface
over the length where the vertical pressure possesses a positive value. For
steady-state motions the deflections and shear forces at a certain x-coordi-
nate do not change in time (Fig. 9.5.32a). Consequently, the integral of
shear forces encountered by one element when running through the con
tact patch produces the total forces and moment acting from ground to
tire.
Under the influence of longitudinal and lateral slip the element under
goes a complex variation of deflections. A /i-slip curve with negative slope
causes the motion to become unstable when internal damping is in
sufficient. In practice, this is reflected in the often observed slip-stick phe
nomenon.
The base point of the elastic element is the point of attachment to the
carcass. It moves in a plane parallel to the wheel center plane. This plane
may be offset with respect to the wheel center plane by the amount v,. rep
resenting a lateral uniform deflection of the carcass due to a lateral force.
Moreover, a longitudinal shift u, of the contact patch may occur due to a
driving or braking force. An angular distortion ft will be disregarded here.
With a rotational speed about the vertical axes ωz and slip speed com
ponents of the lower portion of the wheel with respect to the road, Vcx and
Vcy> positioned at coordinate x and with deflection components u and v:
y. - K« - (vc + vX + u
826 MECHANICS OF PNEUMATIC TIRES

Fig. 9.5.32a. Tread element of "brush type" tire model moving through contact patch.

Figure 9.5.32 illustrates the situation. For motions with relatively long
wavelengths the terms with <o, may be neglected.
The slip speeds are dictated by the values of slip angle, α, and longitudi
nal (brake or drive) slip s. We have:

- - V sin a.

Fig. 9.5.32b. Traction field of brush type tire model generated with hybrid computer. K-fl.5
V~ Vo - 30 m/s. Notation: Ff - fJW. F? - FJW, Hi, - M,/2aW.
ANALYSIS OF TIRE PROPERTIES 827

FIGURE 9.5. 32c. Traction field of tire with flexible carcass moving on surface with sharply
decaying p-slip curve. Notation: Ft - F,/W. F, - FJW, A?, - M,/2aW. For the left-hand
figure y - 0.5 V0 with V, - 30 m/s.

The rolling speed V, with which the element moves through the contact
patch is:

with £2 the rotational speed about the wheel spin-axis and re the effective
rolling radius at free rolling. Instead of re one could take r1 as has been
done in eq (9.5.63). This, however, is not essential when near free rolling
conditions are not the object of investigation.
The i-th passage of the element begins at the instant ti-1 and ends a time
interval Δt later at ti. The interval equals

Coordinate x of the base of the element to be followed is determined by:


x = a - V,(t - /_,).
The element has stiffnesses cx and cy and damping coefficients kx and A:
in x and v direction respectively. At the tip of the element a concentrated
mass, m, is considered to be present. These quantities are taken per unit of
length of circumference. The same holds for the shear force components
qx and qy and the vertical pressure distribution qz. The following equations
govern the motion of the element:
mii + k,ii + c,u — q,
mv + kv + cv — q.
828 MECHANICS OF PNEUMATIC TIRES

When the resulting shear force opposes the sliding speed, we have:

with
V
' s = V
JV' fx 2 +
' V
~ gy

q = M>
and the functions

The vertical pressure distributions may be taken as a function of the verti


cal load W.
The forces and the moment acting on the carcass are equal to the in
tegral of the contributions of the internal forces per unit length

crv.
The respective integrals determining longitudinal force, lateral force and
aligning torque read:

,'dx-V, qy'dt
'«_,

M,~Vr! " {(x + *c + u)qy - (vc + v)?x} dt.


*>l-\

The initial value of the deflection and the time rate of change of the de
flection of the tread element at / = ti 1 are assumed to be equal to zero.
The effective average arcass deflections are denned by the relations:

where C« and Ccy are the carcass stiffnesses and fx and f, the effective frac
tions which take into account the longitudinal and lateral rolling deforma
tion of tire with respect to rim and road (displacement of vertical pressure
center). The values of Fx, Fy and Mz are available at the end of each pas
sage. The values of uc and vc are calculated with the Fx and F,, values ob
tained at the end of the previous passage.
A set of results obtained with the aid of computer circuits representing
the above equations are presented in fig. 9.5.32b. Curves obtained with a
more advanced model exhibiting carcass flexibility represented by Green
functions, for which we refer to the original text, show a close resem
blance. An interesting field obtained with the latter model is shown in fig.
9.5.32c, and may represent tire force generation on icy surfaces (cf. Weber
829

[69]). An extension of the theory has been given by Koch [72] also cov
ering transient and periodic side slip motions.
Approximate Mathematical Representation of Combined Effects
In theories of steady state or quasi-steady state automobile motions, tire
forces and moments must be introduced in an appropriate mathematical
form. The complexity of this representation depends upon the object of
the investigation. Many authors [29, 30, 31, 32, 33] restrict the motions to
be investigated to relatively small deviations from the rectilinear path. In
that case the equations of motion can be linearized and only coefficients
like cornering stiffnesses are of importance. When, in addition, longitudi
nal forces are not taken into account these coefficients depend only on the
vertical wheel load, which may vary due to lateral load transfer. In the lin
ear representation the latter effect is only of importance when initial steer
or camber angles of the wheels front or rear are present. The change in
rolling resistance (cf. part 8.2, fig. 9.2.56 and sec. 9.4.1) with wheel load
will always enter the problem as soon as a finite height of the vehicle cen
ter of gravity is considered [31]. The order of magnitude of the cornering
stiffness CV,, (force per radian), expressed in terms of the nominal vertical
load, lies between 5 W0 and 10 W0; the pneumatic trail t,, expressed in terms
of half the contact length lies in the range 0.4a to 0.7a.
Theories which consider high lateral (cornering) accelerations (e.g. [34]
to [42] and [58, 61]) need a more or less complete mathematical descrip
tion of tire behavior. In these investigations most authors describe tire
characteristics by means of simple mathematical expressions (parabola,
sine, exponential) which correspond more or less to actual characteristics
known from the literature. More sophisticated theories e.g. [38, 42, 58, 59]
show the employment of tire characteristics in which the influence of ver
tical and longitudinal forces are also taken into account.
A complete, more or less exact, mathematical representation of mea
sured data is difficult to accomplish. Fiala has combined his theories for
side slip (with influence of Wand μ) and camber. He presents explicit ex
pressions for Fy and Mz for which we refer to the original paper [13].
We shall present a procedure with which the combination of most ef
fects treated in this part 9.5 can be approximately represented. The prin
ciple of this method has been given in [18]. The philosophy is as follows.
Consider the tire characteristics (Fy vs. α, Mz vs. α) measured at nominal
wheel load W0 and zero camber and longitudinal force (except small roll
ing resistance force). From these basic tire characteristics we attempt, us
ing the theoretical and experimental experience obtained, to derive Fy and
A/, in cases where the conditions differ from basic conditions. These basic
conditions are:

Under these circumstances we obtain


F, - F>0, M, - MJM, CF. = CFm, CH. = €„„.
The functions F^a) and Mzo(a) may be approximated by antisymmetrical
mathematical functions.
The influence of a variation of the vertical load Whas two effects. First
the cornering stiffnesses vary according to functions which may be mea
sured (cf. fig. 9.5.6, low values of a). Secondly the maximum Fy changes
proportionally with iiW (experiments of Borgmann [16] show that in fact
MECHANICS OF PNEUMATIC TIRES

this variation is somewhat less than proportional). Through this effect the
influence of /i has been taken into account. It is assumed that μ effects only
the more or less horizontal level of the cornering force characteristic. On
slippery roads and sometimes to a lesser degree on dry roads, cf. [76], the
shape of the curves may change considerably (fig. 8.3.34) and a different
approach is needed.
The following equations hold approximately when only W and \i differ
from the basic conditions:

(9.5.90)

with the equivalent slip angles

•a. (9.5.91)

The cornering stiffnesses CFα and CMα are assumed to depend only on W,
and are independent of jii. The functions CFα( W) and CMα( W) may be ap
proximated by polynomials. In figure 9.5.33 the transformation of curves
(0) (=hasic) to curves (2), obtained by means of multiplication as in
dicated, result in the characteristics relevant for the new situation (/*, W).
The influence of camber can be approximated by shifting the curves (2)
horizontally, so that for α = 0 the force Fy and the moment Mz equal CFγγ
and CMγγ respectively. For experimental verification cf. Fonda [3, fig. 73],
or Henker [6, fig. 98] and figure 9.3.40. Equations (9.5.94) can be applied
but with different arguments.

«F«, =

(9.5.92)
«*

The functions CFγ( W) and C*,T( W) may be approximated by 1 inear func


tions of W. The quantity CMγ is relatively small and might very well be ne
glected. The new curves (3) are shown in figure 9.5.33a.
The influence of Fx is complicated. We propose the application of the
following formulations which are expected to give reasonable results.
With consideration of the theory of section 9.5.1 the cornering stiffnesses
may be represented by:

V^w) (9.5.93)
1- W, Ff)
ANALYSIS OF TIRE PROPERTIES 831

FIGURE 9.5.33a. Illustration of characteristics obtained from the basic curves (0) according
to formulae (9.5.90-95).
The following successive steps are carried out:
Diagram a: 3 : new conditions (;i. W, y)
3-4 : hor. mult. CrJ(W)/CFJ^W, F,)
0 : basic W Wo. y - f, - 0) 4-5 : red. mult. V - (.F
0-1 : horizontal mult, with Cfa0/C^(W) 5 : new conditions (/i,
1-2 : radial mult, with pW/poWb
2 : new conditions (/i, W)

Diagram c: Diagram d:

0 : basic (p«. Wo, y - F, - 0) 2-3 : horizontal shift


3 : new conditions (p. W, y)
0- 1 : rad. mult, with —— —
Mo W0 3-4: rad. m.
W, F,)
1-2 : vertical mult. CMai
2 : new conditions (/i, W) 4-5 : vert. mult. (1 -

Diagram b: 5-6: vert. add. -£-


I <-<y <-««
2-3 : horizontal shift 6 : new conditions (ft, W, y, F,)

in which C, , and C, Y. denote the lateral and longitudinal stiffnesses of the


carcass relative to the wheel rim. The constant £,, and ζx represent the per
centages of effectiveness of the carcass elasticity. They may be found ex
perimentally or by curve fitting of the whole function CMα (/i, W, Fx).
The approximate formulations (9.5.93) in which the exponents are of
the order of n ≃ 10-20 and m ≃ 1-2, do not take into account the theoreti
cally exact vertical portions (dotted in fig. 9.5.31) where complete sliding
occurs at |F,| = \iW. In order to. describe the variation of CFα and CMα in
832 MECHANICS OF PNEUMATIC TIRES

-.5 BRAKING DRIVING .5 rl/W


FIGURE 9.5.336. Interaction of longitudinal and lateral wheel slip in the generation of
horizontal tire forces, as calculatedfrom a basic force-slip relationship.

these portions, the longitudinal slip ratio actually should be introduced.


We approximate this with a steep approach of the endpoints (1,0) and
(-1, 1) in the diagram (CFα/μW\s. FJpW). Most authors [43, 41] approx
imate CFα (μ, W, Fx) by an ellipse. As a result of this, the curves (Fy vs. α)
are merely multiplied in the vertical direction. This does not agree with
theoretical findings (fig. 9.5.30c). A multiplication of the basic curve
/>(α), mainly in the radial direction from the origin, is expected to occur.
We obtain the following formulae:

w
"1 CMa(W)
M,

Js. (9.5.94)
C,

in which the equivalent slip angles read:


Cf,Qi. W, Ff)
(9.5.95)

Figure 9.5.33 shows the change of the curves due to the introduction of the
longitudinal force Fx. For F^ curve (5) is the final curve and for Mn curve
(6) is the final curve, covering all effects of deviations from the basic con
ditions.
ANALYSIS OF TIRE PROPERTIES 833

It should be pointed out finally that for the description of the unrest
ricted motion of a vehicle, tire characteristics must be used with the ab
scissa sin α, instead of α, extending from — 1 to + 1 (cf. [40]).
These final characteristics are ready for use in the theory of vehicle dy
namics. It should be noted that the formulations above are but approxi
mations of measured behavior. The basic curve measured under nominal
conditions (average vertical load, zero camber, free rolling, actual road
surface) depends on tire construction and inflation pressure. Quantitative
experimental data is needed for the execution of each intermediate step.
The simulation of wet traction behavior is more complicated. In that case
the influence of the speed of travel is appreciable. Due to the fact that the
a curves in the Fx - Fy plane show double valued relations, we must em
ploy an independent variable which changes monotonously along the a
curve. The longitudinal force Fx is no longer suitable as an input variable.
Instead, we may use the longitudinal slip.
A possible approach has been presented in [58]. This study starts with
the introduction of a slip vector with components VCJ\ V\ and Vy/\ V\.
Here, Vcx and Vcy denote the components of the slip speed of a point at
tached to the wheel rim at a distance re (the effective rolling radius at free
rolling, fig. 9.4.10) below the wheel axis. The absolute value of this slip
vector | VJ V\ is used as the abscissa of the basic tire force characteristic
relating the resulting horizontal force to the total slip. In this characteristic
a decay may occur at higher values of slip which is typical on wet roads
and which may also be encountered, although to a lesser degree, on dry
surfaces (cf. e.g. Segel [76]). From the assumption that the resulting force
acts in the opposite direction of Va the force components Fx and F,, are ob
tained. In cases where this assumption does not always hold, an artificial
correction may be given as outlined in the original paper. Figure 9.5.33fe
shows an example of a simulated plot. Instead of the slip quantities men
tioned above, we have used here the more common slip angle α(= —arctan
Vcyl Fcos α) and longitudinal slip s (= Vcx Fcos α) from which Vcx, Vcy and
Vr (= V cos a - Vcx) follow.
Alternative models and methods aimed at the simulation of tire charac
teristics are described in [75 VoLII].

9.5.2 Nonsteady State Motions


In this section two. kinds of nonsteady state motion will be discussed.
First, the response to nonsteady state horizontal motions of the wheel axle
and secondly, the influence of vertical oscillations of the wheel axle upon
cornering force. The wheel plane is assumed to remain in a vertical posi
tion normal to the road. In [66] Segel made an attempt to describe the re
sponse to wheel camber oscillations in an investigation into motorcycle
dynamics.
Horizontal Motions (Transient and Shimmy)
In the theories describing the horizontal nonsteady state behavior of
tires one can identify two trends of theoretical development. One group of
authors assumes a bending stiffness of the carcass and the other bases its
theory on the string concept.
In principle, the string theory is simpler than the beam theory, since
with the string model the deflection of the foremost point alone determines
the path of the tread for certain wheel movements, whereas with the beam
834 MECHANICS OF PNEUMATIC TIRES

model the slope at the foremost point also has to be taken into account as
an additional variable. The latter leads to an increase in order of the sys
tem by one.
Probably the first investigator who tried to describe tire behavior
mathematically in the study of shimmy is Kantrowitz [44] (1937). In spite
of his rough and theoretically unsatisfactory assumptions, the theory de
veloped gave a fair correspondence with measured values of divergence of
wheel deflections and frequency of the shimmy motion. Kantrowitz stud
ied the damping effect of the gyroscopic couple due to lateral distortion of
the rotating tire. Another theory, apparently inspired by Kantrowitz'
work, was developed in 1942 by Greidanus [45]. Where Kantrowitz' work
shows features of both the beam and the string, Greidanus is consistent in
applying the bending concept in his interesting study. Besides the slope,
the curvature of the peripheral line just in front of the contact point is also
important for the further development of the motion. In Greidanus' model
a vanishing area of contact was considered as may be deduced from his
approach. In a discussion on Saito's paper [46] Pacejka has given the dif
ferential equations which govern the kinematieal variations in lateral tire
distortion for the beam type model with finite contact length. These equa
tions appear to be identical to those given by Greidanus when the influ
ence of camber is not considered, and when the contact length is taken
equal to zero.
In 1962 Saito [46] presented a theory using a tire model consisting of an
elastic beam of which a finite length makes contact with the road. The the
ory is based on an approximate treatment of the kinematic behavior of the
contact line. Frequency response curves are given for the force and mo
ment with respect to lateral and angular motions of the wheel plane. In or
der to obtain better agreement with experimental results, Saito introduced
theoretically unjustifiable empirical corrections.
Besides this group of investigators which were inspired by the work of
Kantrowitz, another group exists which has studied the problem with the
aid of tire models more or less based on the string concept.
In 1941 Fromm [47] gave a simple theory where this model (although
not mentioned by name) is investigated for the case of point contact. A
similar theory was developed by Bourcier de Carbon [48] in 1948 together
with an extension, somewhat unclear, which increases the order of the sys
tem by one. A similar simple theory originates from Böhm [49]. He uses
the nonlinear steady state cornering characteristics in order to find the am
plitude of the periodic shimmy motion.
In 1941 von Schlippe [25] presented his well-known theory of the kine
matics of a rolling tire, and introduced the concept of the stretched string
model. For the first time a finite contact length was considered. In the
same paper Dietrich applied this theory to the shimmy problem. Mathe
matical difficulties arose in the form of transcendental equations, due to
the retardation effect of the assumption of a finite contact length. Later on.
two papers of von Schlippe and Dietrich [50, 51] were published in which
the effect of the width of the contact area is also considered. Two rigidly
connected coaxial wheels, both approximated by a one-dimensional string
model, are considered. The strings and their elastic supports are also sup
posed to be elastic in the circumferential direction.
Segel [52] derived the frequency response characteristics for the one-di
mensional string model, and these are similar to response curves which
arise in Saito's approximate theory for the beam model.
ANALYSIS OF TIRE PROPERTIES 835

Smiley [10] gave a summary theory resembling the one-dimensional


theory of von Schlippe [25]. He has correlated various known theories
with several systematic approximations to his summary theory.
In [19] Pacejka gave the nonsteady state response of the string model of
finite width provided with tread elements. The important gyroscopic effect
has been introduced and the nonlinear behavior of the tire due to partial
sliding has been discussed. Applications of the tire theory to the shimmy
motion of automobiles have been presented. In [64] the effect of mass of
the tire has been investigated with the aid of an exact analysis of the be
haviour of a rolling stretched string tire model provided with mass. This
complicated and cumbersome theory, not suitable for dynamic studies,
was then followed by an approximate, more convenient, theory [65] taking
into account the inertial forces only up to the first harmonic of its distribu
tion along the tire circumference. An outline of the theory together with
experimental results will be given in a subsequent section.
Rogers derived empirical differential equations [62] which later were
given a theoretical basis [63]. As a result of Rogers' research, shimmy re
sponse of tires in the low frequency range (mass effect not included!) can
be described satisfactorily up to rather high reduced frequencies (i.e. low
wave lengths) by using simple second-order differential equations. Sekula
et. al. [68] derived transfer functions from random slip angle input test
data in the range of 0.05 to 4.0 Hz. From this information cornering force
responses were deduced for both radial and bias-ply tires to slip angle step
inputs.
Ho and Hall [67] conducted an impressive experimental investigation
using relatively small aircraft tires tested on a 120 inch research road
wheel up to an oscillating yaw frequency of 3 Hz. A critical correlation
study with theoretical results revealed that reasonable or good fit of the ex
perimental frequency response plots can be achieved by using the theoreti
cal functions (9.5.1 14, 122, 123) presented in the next section. It should be
pointed out that in testing small scale tires certain similarity rules, enun
ciated in one of the following sections, should be obeyed.
Transfer functions
We shall discuss here some theories of nonsteady state tire motion,
based on the stretched string concept, starting with the relatively simple
case where sliding in the contact area does not occur (/i -» oo or a -» 0 and
4> —» 0). The response of the force Fy and the moment Mz with respect to
arbitrary variations of the slip angle a and the spin <£ will be determined
for models which are successive approximations of the stretched string
model with tread rubber elements shown in figure 9.5.9, of which the
steady state behavior has been treated in section 9.5.1.
The contact equations (9.5.9) apply when the velocities of sliding Va
and Vsy are taken equal to zero. They read then:
du_ _ d_u_
dx ds
(9.5.96)
r~ — 7- = —a 4- x&.
dx ds
These partial differential equations will be solved by using Laplace trans
formation. The Laplace transforms have been written in capitals. We will
836 MECHANICS OF PNEUMATIC TIRES

not transform with respect to time, as is done usually, but with respect to
the distance travelled s = Vt, where V is a constant. The Laplace trans
form of a variable quantity, generally indicated by q, is denned through:

/ e-»q(s)ds.
Q(P.) = »0 (9.5.97)

With the initial condition u(x, 0) = v(x, 0) = 0 at s = 0 we obtain:


dU
(9.5.98a)

(9.5.986)

The solutions of these ordinary first-order differential equations read:

U = C.je"-' + - >>*, (9.5.99a)


P,

V = Cje'-* + - 1- + x\ $. (9.5.996)
P' \P' I

The terms Cue"-* and C,e""* point to a retardational behavior. The coeffi
cients Cu and C, are constants of integration. They are functions of p, and
depend on the tire construction, expressed for example by equation
(9.5.37) and the boundary condition (eqs. (9.5.39) and further). The condi
tions at the leading edge are:
x = a: u = 0orl/ = 0 (9.5.100)
leads to the following expression for C.:
1
(9.5.101)
t
For the determination of C, we turn to eq. (9.5.37) whose transform is

(9.5.102)

With eq. (9.5.996) the following solution is obtained:


Vc - C+e'"< + C-e-"°< + (1 - e2) I , C*" a
[ 1 - OcP,

<t\. (9.5.103)
P, P, P,
The three constants of integration C, C and C_ may be solved with the
aid of boundary conditions discussed before (cf. eq (9.5.39) and further).
ANALYSIS OF TIRE PROPERTIES 837

With the use of eqs (9.5.26-28) and (9.5.40) the Laplace transform of
lateral force and moment can be obtained. For more details we refer to
reference [19]. The discussion will be continued for the simpler model
without lateral flexibility of the tread rubber elements (cp —» oo, ν = νc).
The longitudinal flexibility of the tread elements will be maintained. The
constant of integration Cν appearing in (9.5.99b) can now be found with
the aid of the condition at x = a expressed by (9.5.39), with ν = v... Ex
pression (9.5.99A) then becomes:

p. p. *
(9.5.104)
At the leading edge the deflection becomes:
a

or transformed back:

-ji + — = a- a(j> = ^- —— <z -j-, (9.5.106)


ds a ds ds

where use has been made of the relations (9.5.2-4).


The first-order differential equation may be found immediately from
the original differential equation (9.5.10), when the condition that for x =
a the slope becomes dv/dx = — v,/a is taken into account.
For the calculation of the lateral force Fy and the moment Mz acting on
the simple string model, two methods are encountered in the literature.
The first method, employed by von Schlippe, makes use of integration of
the internal lateral forces along the length of the string extending from
minus to plus infinity. For the calculation of the moment, von Schlippe in
troduced a correction factor (ρ in [5 1, eq (77)]) with which the influence of
the circular tire of radius r is meant to be expressed. It turns out that the
effect of the circular tire upon the moment about the vertical z-axis
through the contact center generated by the internal lateral forces ccν is
completely cancelled out by the torque exerted by radial reactive forces
S/r produced by tension in the string which is stretched around an imagi
nary cylinder and shifted laterally a distance v. Smiley [10] and Hadekel
[9] adopted the same erroneous correction factor. Consequently, the circu
lar shape of the tire (string) has no effect (except a small influence due to
the finite length of the circular string (cf. Frank [12]) and the string can be
considered to be developed in a plane (i.e., the road surface).
The other method which leads to the same results is due to Temple [9]
who integrated the internal lateral forces only over the contact length and
added the influence of the internal tension forces in the string just outside
the contact region.
According to Temple's method we obtain for the lateral force:

vdx + S(v, + vz)/a (9.5.107)


838 MECHANICS OF PNEUMATIC TIRES

and for the moment due to lateral deformations denoted by M'/.


i

= cc ] vxdx + S(a + a) (v, - v2)/a (9.5. 108)


* -a

where ν1 and ν2 are the deflections at x = a and -a respectively, and S =


σ2cc according to eq (9.5.34).
The moment due to longitudinal deformations of the tread elements u
denoted by M* becomes with eqs (9.5.26) and (9.5.40):

M* = -cfx I I uydxdy. (9.5.109)


t-a 1-b

By adding up the contributions (9.5.108) and (9.5.109) the total moment


about the z-axis is obtained
M, = M', + M*. (9.5.10)
The Laplace transforms of F,,, M', and M* can readily be obtained now
with the aid of eqs (9.5.34), (9.5.104), and (9.5.99a). In general, these
transforms can be written in the form:
F,A
M'ji + A/;* (9.5.111)

in which F,, F+, A/1, A/;, M* and A/* represent the transfer functions of F^
M', and M* with respect to the slip angle α and the spin <f> (=dψ/ds). In or
der to avoid double subscripts, the subscripts y and z are omitted. We find
the following transfer functions in vector form for the tire model consid
ered
c<_ ^
P. P, op,+

(9.5.112)

(A/1, (0, -1)

a(\ + e-*") +/>,{g(q + a) - I/A2) (1 -

X -l,a (9.5.113)

and furthermore
ANALYSIS OF TIRE PROPERTIES 839

in which the quantity has been introduced:


K* - % c?Vc,x. (9.5.115)
By transforming back the above expressions, the deflection, the force and
the moment can be found as a function of distance travelled for given var
iation of α and </>.
Response to step function of the slip angle
An important characteristic aspect of tire behavior is the response of the
lateral force to a stepwise variation of the slip angle α. The initial condi
tions at s = 0 read: ν(x) = 0; for s > 0 the slip angle becomes α = α0. From
(9.5. 104) we obtain for the lateral deflection of the string in the contact re
gion:
v
(for x> a — s) (9.5.116)
«„
while for the original points the following simple relation holds:

—=s (for x < a — s). (9.5.117)

With (9.5.107), finally, the force has been calculated for the two intervals,
with and without original contact points.
(s**2a) (9.5.118a)
Fy = c, {2(o + a)2 - 2a «„ (s > Id). (9.5. 1 1 86)
The latter part (9.5.118b) could have been obtained immediately from
(9.5.1 12). The variation of F^ graphically shown in figure 9.5.34, may be
used for an experimental determination of the relaxation length σ of the
tire. For this purpose the ratio of the force attained at s = 2a and the
steady state value CFαα0 may serve. Another method for the determination
of the relaxation length has been given in section 9.5.1, eq (9.5.51).
Response to sinusoidal inputs (shimmy)
The frequency response functions for F and Mz can easily be found by
replacing />, by iωs in eqs (9.5.1 12-1 14). The path frequency ωs equals 2ir/
A, where λ denotes the wavelength of the motion. When we are dealing
with sinusoidal motions with the x -axis deviating only slightly from the x-
axis fixed to the road (cf. fig. 9.5.1), it is convenient to replace α and <j> by
the variables ψ and y or /8(= dy/ds for ft <K 1). With the aid of the relations
(9.5.2), (9.5.3), and (9.5.8) we find for the transfer functions with respect to
•y and /•! (or v), expressed in terms of the transfer functions found before:

(9.5.119)

j_
' P,
840 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5.34. The response of the lateral force F, on a step input of the slip angle a,
calculatedfor the relaxation length a = 3a.

The frequency response functions, F+(iu,) etc., are the complex ratios be
tween output, F\ etc., and the input ψ etc. For the tire model with and
without tread elements their absolute and phase relationship has been cal
culated. For shimmy analysis, the response of Mz to ψ is of greatest impor
tance.
In figures 9.5.35a, b, c, d the various responses are shown, as a ratio to
their steady state values as a function of the nondimensional path fre
quency, ωsa. Figure 9.5.35e shows the response M* divided by the constant
κ*/a, the value of which A/J approaches when «, —» oo. The approximate
responses treated below are also shown in these figures. The phase angles
<j> are taken as positive when the output lags behind the input, which ap
pears to be the case with the force and moment due to lateral deformation.
The moment M* due to longitudinal deformation however, appears to
lead in phase. The phase lead of M* causes a reduction in phase lag of the
total moment Mz with respect to ψ, as has been illustrated in the diagram
of figure 9.5.36. This is a favorable effect for the suppression of shimmy.
In the complex plane shown in figure 9.5.36 the response curves are
drawn for the moment M',, which applies for an infinitely thin tire, and for
the moment Mz = M', + M* for a tire of finite width with κ* = CMαa. The
moments are made nondimensional by division by the steady state value
Mzo, which occurs at α = α0 = fy,. The curve for a tire with κ* = CMαa is
obtained by vector addition of M1, and Mz. Curves for other values of κ*
may be obtained by multiplying the vector of Mz by the factor κ*/CMαa.
The calculated behavior of the linear tire model has unmistakeable points
of agreement with motions found experimentally at low values of the
swivel frequency. At higher values of the frequency, the influence of the
841

1 tfej ® Ifel ® 1 ftJ ® 1 Prtl ®


\
fl v
_/ ^
/
s \
\ ^s S
/,
.'
X r- ^K<-
^ ~*^= *
S
N -•<- •—-—.
^*s ~^~^t —
| -5
1
\ v \ V ^^ .
S^ S ik^
^r ~~—,• -— •c->
V1 --- =-
~~,
» ^kx
•f
^"~~ h—
•---. ,
exact with tread elements

1
* . ® yt ^ - exact without tread elements
•' or von Schlippe approx
'.^>
OS ^\ approx. parabola
2 7
pure damping

/ 1 approx. straight tangent


-*> "••« —-—•
~—. t~^. i
±
FIGURE 9.5.35. Exact and approximate response characteristics offorce and moment with
respect to $ and ft for string type models (a* = 3a, a = 3.74a, « — 1/7.5; without profile
elements: a = 3a).

finite
tread
width
z ero i
t read
width

FIGURE 9.5.36. Response curves in the complex planefor the moment M, with respect to yaw
angle \/>.
842 MECHANICS OF PNEUMATIC TIKta

gyroscopic couple due to tire deformation, dealt with later on, is no longer
negligible.
We may note, furthermore, that above a certain value of K* the curve
for the total couple M,/MM will not encircle the origin but will remain on
the right-hand side of this point. This appears to be a typical property of
curves obtained experimentally, which has not been explained before. The
point of intersection of the curve for Mz and the real axis (fig. 9.5.36) rep
resents the point of "kinematic" shimmy (cf. Kantrowitz [44] and Saito
[46]). This sort of shimmy may occur at very low values of speed of travel,
where the frequency and consequently the moments due to viscous damp
ing and inertia acting about the king-pin axis become very small.
Dynamic tire tests at low values of the swivel frequency show good cor
respondence with the theory as will be shown after the introduction of the
gyroscopic effect. Experiments with tires of different tread shapes indicate
that tires with longitudinal ribs have a greater K* than tires with a block
profile. Tests indeed show that ribbed tires are less sensitive to shimmy
[19].
The response for ωs ∞» oo (\ = 0, standing tire) might be used for experi
mental determination of parameter K*. Torsion of a standing tire about the
vertical axis over a small angle i/*,, produces a moment:
-M, = Ok = (CHa + KVa)fc (9.5. 120)
from which the quantity K* can be obtained after the determination of the
aligning torque stiffness.
Figure 9.5.35 shows that the exact responses of the models with and
without tread elements are qualitatively the same. Quantitatively, the de
viation from the response of the model with tread rubber becomes larger
for shorter wavelengths (larger ωs). In the region important for shimmy
analysis, at the left-hand side of the hatched band, very good correspon
dence exists. The exact treatment of the simpler model, however, is still
too complicated to be used in the actual shimmy analysis. We shall there
fore consider three approximations for the response of the model without
lateral tread rubber flexibility.
Approximations
A first approximate description of the behavior of this model originates
from von Schlippe [25]. The contact line is considered to be a straight line
connecting the two endpoints of the real contact line (see fig. 9.5.37). Only
the deflections ν1 and ν2 of these points are of importance now. For the
transformed force and moment we obtain, after some manipulation and
with the aid of expressions (9.5.50):

cc(o + a) ( V, + V2) = CFa ~,

ce (0(0 + a) + V,a*} ( V, -

With the aid of (9.5.104) the transfer functions become:

(9.5.122)
ANALYSIS OF TIRE PROPERTIES 843

(JVC A/;) , -1) + (-1, a + a + (9.5.123)

The responses with respect to ty and ft, obtained with the aid of formulae
(9.5.119), appear to coincide practically with those obtained from for
mulae (9.5.112) and (9.5.113) in the range of wavelengths investigated for
σ = 3a (see curves "exact/v. Schlippe" in figs. 9.5.35-36).
Simulation of the von Schlippe representation by means of an analog
computer appeared feasible although complicated (cf. [19]). In this simu
lation use has to be made of equation (9.5.106) for obtaining ν1, and of the
retardational behavior in order to generate ν2. The latter may be carried
out with the aid of a memory device (magnetic tape recorder or other
wise).
Simpler directly applicable approximations are obtained by expanding
the exact response functions (9.5.112), (9.5.113) and also (9.5.114) into
powers of ps. With the use of relations (9.5.1 19) we find for the power se
ries of the responses to the angular displacement ;// and the lateral dis
placement y of the wheel plane:
Ft(ap, + 1) = CFa(l - ap,) + 1crf(a + a) ((a + Ka)p]

F^pp, + 1) = -CF,P, + Ic^a* (a + ^a)p]

(9.5.124)
M'^ap, + 1) = -CHa(\ - ap,) -(a + a)aCM,p2,
a)a\CMa - O.C

- 0.0667 c^)p]

straight connection
< v Schlippt)

FIGURE 9.5.37. The exact deflection of the simple stretched string model and three
approximations.
844 MECHANICS OF PNEUMATIC TIRES

= -K*(p,
(9.5.125)
A/?=O.
In the periodic case we have: ps = ia, = iω/ V = 2iri/λ. When in the power
series (9.5.124) truncation is made to the second power o(pn the shape of
the contact line is approximated by a parabola and when the second and
higher power are neglected, the contact line is approximated by a straight
line, both touching the real contact line at the leading edge (cf. fig. 9.5.37).
The larger the wavelength becomes relative to the contact length, the bet
ter the approximation will be. In the practical range of wavelengths, espe
cially for the amplitude responses of M', to \j/, the parabolic approximation
furnishes a great improvement with respect to the straight tangent approx
imation. The responses gotten from these approximations are shown in
figure 9.5.35. Also, the response due to longitudinal deformations (9.5.125)
has been approximated according to a quadratic and a linear variation of
these deflections along the x-axis (see fig. 9.5.35e). The linear representa
tion (up to /;,) corresponds to a viscous damping. The approximate differ
ential equations for the force and the moment, directly applicable for
shimmy analyses, which correspond to the quadratic approximation of the
response functions (parabolic lateral deflection) read:

dy
-

ds > (9.5.126)

JL_ i i.
a)
ds2 ds2

ds

When omitting the underlined portions of eqs (9.5.126), the linear approx
imation (lateral deflection according to straight tangent) is obtained. In
that case Fy and M', only depend on ν1 (cf. eqs (9.5.106), (9.5.46) σ* = σ,
(9.5.49)), and M* corresponds to a moment due to viscous damping
(damping coefficient = κ*/V).
It should be emphasized that to the constants €,,„ and (',„,. values might
be given obtained from steady-state full-scale slip experiments. These val
ues, particularly their quotient, do not necessarily agree with theoretical
evidence from the simple string concept.
A final approximation can be obtained by neglecting the dimensions of
the contact area (a = b = 0). This leads to a representation of the non-
steady state tire behavior which has been used by various authors. Al
though the moment should vanish for the string model, some authors
maintain M′ and use the measured aligning torque stiffness ('„„.
ANALYSIS OF TIRE PROPERTIES 845

Nonlinear behavior due to partial sliding


The general treatment of the combination of nonsteady state drift and
turning is extremely difficult. We shall therefore restrict ourselves to the
simple string type tire model without lateral tread rubber elasticity. The
effect of tread width (A/?) may be treated separately wh'en longitudinal
sliding velocities are assumed to be small. From the above analysis it ap
pears that for a wheel swivelling about a vertical axis through the wheel
center, the two moments M* and M′ do not attain their maximum at the
same time. A phase difference ranging from 1/4π to about Viir radians will
occur in the practical range of wavelengths. In this range of wavelengths,
many times larger than the contact length (order of 10 times), the shape of
the deflection of the tire approaches the stationary shape, i.e., shows an al
most straight contact line in case of complete adhesion (linear approxima
tion mentioned above). If, furthermore, in figure 9.5.15 the front sliding
region is neglected, the lateral deflection at the leading edge v, is the only
quantity which governs the lateral tire deformation. As long as no total
sliding occurs, ν1 varies according to the differential equation (9.5.106). Fi
nally, for Fy and M'., measured or calculated characteristics as a function
of slip angle α = ν1/σ may be employed. According to the steady state side
slip theory, the relaxation length of a tire with tread elements decreases
somewhat with increasing slip angle (fig. 9.5.11). The introduction of a
nonconstant a, however, would lead to great mathematical complexity.
In [19] Pacejka uses an analog computer in connection with a tape re
corder for the simulation of a more exact representation of the non-
stationary behavior of the partially sliding tire, which can be seen as a
nonlinear extension of von Schlippe's approximation of the behavior at
complete adhesion.
Gyroscopic couple due to lateral deflection of the tire
In [19] it has been shown that inertial effects due to lateral vibrations of
the peripheral line of the nonrolling tire may be neglected for frequencies
lower than about 8 Hz. At higher frequencies tire inertial forces are ex
pected to have considerable influence upon the tire deformation and con
sequently upon tire force and moment response. Approximate calculations
indicate that the ratio of inertial and elastic forces amounts to 6 percent
and 24 percent for a frequency of 7 c.p.s. and 14 c.p.s. respectively.
In case of a rolling tire we have to deal with the substantial acceleration
of a material point of the tire. For a point P (fig. 9.5.38) of a tire with an
gular speed SI and radius r we obtain:

(9.5.127)'
dt2 d? dxdt Bx1
The first term represents the inertial force discussed above. We shall ne
glect this term in the following. When calculating the moment about the
vertical axis, taking into account the last two terms of (9.5.127) together
with the centrifugal force acting on a tire element in radial direction, and
furthermore a lateral deflection ν(x) of a massless tire model according to
eqs (9.5.38) and (9.5.104), we find that only one term remains, which rep
resents the gyroscopic couple. This couple corresponds to the gyroscopic
couple which would arise when the circular peripheral line of the tire is
tilted about the horizontal line which lies in the wheel plane and passes
M6 MECHANICS OF PNEUMATIC TIRES

through the wheel axis. The lateral deflection of the tilted peripheral line
represents the first odd harmonic of the Fourier expansion of the lateral
tire deflection. This imaginary angle of tilt (-γt) will be approximately
proportional to the lateral tire force Fy. We obtain, then, for the gyroscopic
couple:

-Ctyry, (9.5.128)

in which /; represents the effective fraction of the tire polar moment of in


ertia. Cgyr is a tire constant which is proportional to tire mass, mt and tire
lateral compliance, 1/Cy, and is furthermore influenced by the type of tire
construction expressed by the dimensionless parameter c,s::
C^-c^mJCr (9.5.129)
For a conventional tire we found that r = 0.12 is a value which produces
a good approximation to experimentally obtained responses.
From (9.5.128) we obtain the following expressions for the transfer-
functions of A/gyr

(9.5.130)

from which we see that for a certain wavelength \ of the swivel motion,
where ps = iωs = 2iri/λ is a constant, the response increases quadrat ically
with the speed of travel V. In the complex plane, the vector of M^, will be
directed perpendicular to the vector of Fy. In figure 9.5.39a theoretically
obtained response curves of the moment with respect to ψ are shown for
the model with lateral tread rubber elasticity (σ* = 2.4 a, e = 1/7.5). The
value of σ* has been experimentally determined with the aid of eq (9.5.51)
for a tire whose experimental response curves are shown in figure 9.5.39b.
The value of e has been estimated. Figures 9.5.35-36 show that the use of
the simpler model without lateral tread rubber flexibility (exact or para
bolic approximation) is expected to yield results which are close to the the
oretical curves of figure 9.5.39 in the practical range of the dimensionless
path frequency ωsa < 0.35. The straight tangent approximation needs
longer wavelengths for acceptable results.
A reasonable correspondence between theory and experiment appears
to exist. The figures clearly show that an important time influence exists.
With increasing swivel frequency the curves of the moment rise, which
means that the phase lag decreases and thus the degree of self-excitation
of the system becomes less.
The points of the experimental curves obtained at V = 0 (standing tire)
are situated above the real axis and become lower and further to the right
at increasing frequencies. This may be due to the viscoelastic properties of
the tire rubber, which shows larger stiffness and less damping at higher
frequencies. The amplitude of the force f\ appears to increase with in
creasing frequency. A more advanced model outlined below is needed to
produce this expansion of the F-curves. For shimmy analysis the effect of
such an increase in amplitude appears to be of less importance.
The machine with which the shimmy response tests have been carried
out is shown in figure 9.5.40. The wheel is swivelled as it rolls over the
847

FIGURE 9.5.38. The deflected tire peripheral line.

drum, together with the whole structure in which the wheel axle is
mounted. The structure is excited against four coil springs in the reso
nance frequency so that only a small force of excitation is needed. A spe
cial measuring hub (described in sec 8.2) has been used for the measure
ment of forces and moments. Because of the fact that the wheel inertia
distorts the signal, a correction is needed in order to obtain the torque Mz
acting from drum to tire. Results of other shimmy response tests can be
found in references [46] and [57].
More advanced theory of the influence of tire mass
Recently, a theoretical model has been developed giving a better repre
sentation of the influence of tire mass (cf. [65]). Besides this more ad
vanced approximate theory, an exact theory of the anti-symmetric fre
quency response of a string-type tire model with mass has been developed
(cf. [64]). Qualitatively, the agreement between results from both theories
is quite reasonable.
A description of the improved approximate theory and discussion of the
resulting response curves in relation to measured data will be given below.
For the governing equations we refer to the original article [65]. The the
ory has been constructed in such a way that the response equations al
ready derived for the massless tire can be used.
For the sake of simplicity it is assumed that the tire mass is uniformly
distributed and concentrated in a ring coinciding with the equatorial line
848 MECHANICS OF PNEUMATIC TIRES

210km/h

o n .1 e/t
— * — n . 2c/§
— •— n mi eft
—• — n .8 e/i

FIGURE 9.5.39a. Influence offrequency of motion upon response (theoretical).


«•- 2.4a, €- 1/7.5. a - 0. 1 15 m, CFa - 64000 N, Cu. - 3900 Mm, C, - 165000 N/m. C^ - 2.5 x 10"5 s2, m, - 35 kg.

Im

— a— n • 1 e/»
— +— n . 2.16 c/i
—•— n . < c/s
—•— n . 8 e/«
<|).»0.75*

FIGURE 9.5.39b. Influence offrequency of motion upon response (experimental).


Tests carried out on drum lest stand (cf. figure 9.5.40). Tire 9.00-16 block tread pi - I.7S bat.
ANALYSIS OF TIRE PROPERTIES 849

«—^

FIGURE 9.5.40. Shimmy excitation tea standfor measuring tire frequency response.

of the tread band. This ring is elastic-ally supported to the wheel-center-


plane (i.e. to the rim).
When the wheel plane is oscillated about the vertical axis and in axial
direction, inert ial forces arise which act upon the elements of this ring in
axial direction. The distribution of these forces along the circumference
will be approximated in that only the average of the inertial forces plus
their first harmonic with wavelength equal to the circumference will be
taken into account. Consequently, higher harmonics of the distribution of
inertial forces will be neglected. This constitutes the essential approxima
tion introduced in the theory. Obviously, the 'dynamic' tire deformation
which is directly caused by this approximate inertial force distribution
varies along the circumference according to an inclined plane.
Beside these inertial forces, we have the external lateral ground force Fy
and moment Mz which also cause the tire to distort. The zeroth and first
harmonic of this 'static' distortion vary according to a linear relation. It is
of interest to note that the internal force and moment acting on the tire
which are due to this deformation are in equilibrium with the externally
applied F and Mz.
Up to the first harmonic we obtain the total linear deformation, which is
the sum of the static and dynamic distortions. With respect to the vertical
(x, z) plane fixed in space, the displacement of the tire plane is defined by
the sum of wheel plane motion and total tire deformation. The inertial
forces which are the result of these displacements of the tire center plane
have a resultant axial force and resulting moments about the x and z-axes.
For a tire model with given inertial and elastic properties we can now
derive six simple equations which relate inertial forces to dynamic dis
placements and external ground forces to static displacements of the tire
center plane. For given input of the wheel center plane the only unknowns
yet to be established are the external ground force and moment.
This lateral force F,. and moment Mz can be obtained approximately by
introducing an effective wheel plane motion equal to the actual wheel
plane motion plus the dynamic tire center plane distortion. With respect to
this effective wheel plane position the tire shows deformations governed
by the external ground forces only. We may find the response of these ex
ternal force and moment with the aid of the known massless non stationary
tire response theory. As the effective input we take the lateral and angular
850 MECHANICS OF PNEUMATIC TIRES

motion of the line of intersection of the effective wheel plane and the road
surface.
In view of the application to vibration problems it is of greater interest
to determine the equivalent force and moment which act from the road
upon the rigid tire. They are obtained by adding the inertial axial force
and the inertial (including the gyrosopic) couple which are caused by the
lateral tire deformations, to the ground force and moment respectively.
The result should correspond to force and moment responses measured in
the wheel hub after a correction has been made for the inertial force and
couple which act on the vibrating rigid wheel plus tire when lifted from
the ground.
Figure 9.5.41a shows the calculated Nyquist curves for the responses of
the equivalent force and moment to swivel motions around the vertical
axis through the wheel center. The massless response has been accounted
for by using the von Schlippe approximation. The curves represent the
simulated response of a conventional cross-ply tire (left diagram) and of a
belted radial ply tire (right diagram).
First of all we observe that the response curves tend to the kinematic
limiting case when the frequency of excitation approaches zero. For
higher frequencies the moment response curves tend to rise which result in
an increase in damping of the system. The upward ends of the curves,
where the speed has the greatest value, are different for the two kinds of
tires. We also see that the force response curves expand with increasing
frequency, which could not be brought about by the simple approximate
dynamic theory discussed above.
The values of masses, moments of inertia and stiffnesses have been de
termined by means of natural frequency tests and by taking into account
an estimated effective fraction (one half to two thirds) of the tire mass. The
cornering and aligning stiffnesses to which the responses tend when the

FIGURE 9.5.41a. Theoreticalfrequency response curves of'force and moment to yaw angle f
simulating measured response shown in next figure. (From [65J).
ANALYSIS OF TIRE PROPERTIES 851

frequency approaches zero ( V constant), have been obtained by means of


steady-state slide-slip tests. Other quantities such as the relaxation length a
and the component of the turning slip stiffness κ* which influence the mas-
sless response, have been found from additional static deformation experi
ments (cf. eqs (9.5.51 and 120)).
The results of the actual shimmy excitation experiments with the cross-
ply tire are shown in fig. 9.5.4 1b. The typical features exhibited by the the
oretical curves are substantiated very well by these experiments. Also in
quantitative respect we note a fair agreement between theory and experi
ment.
Fig. 9.5.41c shows the calculated dimensionless frequency response
functions of the moment and the force to lateral motions (y) of the wheel
plane. These theoretical functions could not be verified due to lack of ex
perimental data. We note, however, agreement with curves resulting from
the "exact" theory discussed in [64].
The approximate theory appears to be particularly suitable for investi
gations in the frequency domain. In the time domain instability of the tire
model may appear which can be suppressed by employing, for instance,
the von Schlippe approximation for the massless tire response.
Similarity rules for testing with small scale model tires
The equations governing the dynamic performance of tires may be writ
ten in dimensionless form. In these equations, dimensionless quantities ap
pear which should be kept constant in order to ensure similarity between
real and experimental conditions. In the first place, geometric similarity
must exist, that is, ratios of lengths must remain unchanged. Then, the ra
tio between inflation pressure pi and elastic moduli must be kept constant
in order to ensure stiffness similarity. In addition, the following conditions

RADIAL- PLY STEEL BELT

FIGURE 9.S.41b. Experimental force and moment response curves for cross-ply and belted
radial-ply passenger car tires according to experiments on drum test stand. (From [65]).
852 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5.4 lc. Theoreticalfrequency response curves offeree and moment to lateral wheel
plane motion y. Same parameters as in Fig. 9.5.41a. (From [65]).

must be obeyed to maintain kinematic and dynamic similarity respec


tively:
wr «V
— = constant, - constant (9.5.131)
" Pi
where ρ denotes the mass density of the tire. The kinematic condition sim
ply states that the wavelength of the path described on the road must be
scaled down at the same rate as the other length parameters. The dynamic
condition implies that the excitation frequency regarded in terms of natu
ral frequency of full scale and model tire must remain the same. At fixed
ρ/pi a reduction of radius r would require an increase in excitation fre
quency ω at unchanged speed V. Ifpi can be reduced without seriously vi
olating the stiffness similarity (that is, if pi is the main source of tire stiff
ness) V and co may be kept in an acceptable range of operation.
Effect of a Time- Varying Load
Due to road roughness, variations in vertical tire load occur which cause
unfavorable changes in the tire cornering behavior. This has been found
experimentally by, amongst others, Kurz [53]. The long wavelength phe
nomenon can be explained sufficiently by consideration of the steady state
tire characteristics. As has been indicated by Kurz, the static loss of the av
erage cornering force is due to the typical nonlinear variation of cornering
force with normal load (cf. fig. 9.5.42). Tests carried out by Metcalf [55]
(cf. also Endres [54]) reveal that short wave length motions show an addi
tional loss in side force, which Metcalf has termed the dynamic loss, al
though the cause may be entirely of a kinematic nature.
ANALYSIS OF TIRE PROPERTIES 853

FIGURE 9.5.42. Effect of slip angle and normal load on tire side force (from Kurz [53]).

In [56] Pacejka made mention of a theoretical study dealing with this


problem. The essence of this theory, which has been restricted to small slip
angles, will be enunciated below. After that, a semi-empirical theory due
to Metcalf will be given, and finally the direct interaction of vertical and
horizontal tire motions will be treated approximately according to a the
ory of Böhm.
Vanishing values of slip angle
For the theoretical explanation of the behavior of a tire drifting over a
rough road, theoretical tire models will be employed of which the steady
state and shimmy properties have been dealt with in sections 9.5.1 and
9.5.2. The more advanced string tire model with tread rubber elements ap
pears to be the most promising in predicting the real response. The analy
sis is restricted to the case of complete adhesion in the contact zone, so
that in fact the variation of cornering stiffness at zero slip angle will be de
termined for the model considered.
In the theory below, the slip angle a and the direction of motion of the
wheel axle remain constant. Camber and longitudinal forces are not con
sidered. Figure 9.5.43 shows the model in an arbitrary situation. When the
contact center is moved a distance A.V, the following changes occur simul
taneously: lateral shift of wheel plane, forward rolling and change in con
tact length. These changes influence the position of the leading and trail
ing edges of the string.
The x-coordinates of these points read:
' s+ a
>s — a
of which the changes become
Ax, = AJ + Aa
(9.5.132)
Ax2 - A* - Aa.
854 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5.43. Two successive positions of tire model with varying contact length rolling at
constant slip angle a (sliding not considered).

The changes in front and rear lateral deflection v,, 2 are composed of con
tributions due to various causes. For the deflection at the front we obtain:
1. due to lateral wheel displacement (oA,):
Avcl = A(d)
2. due to loss of contact rear (Ax2 > 0):

Avcl = -

3. due to loss of contact front (Ax, < 0): (9.5.133)


Avcl = B*(d) "' ~ Vcl Ax,
"c

4. due to longitudinal displacement of leading edge (Ax,):

Similar expressions are obtained for the contributions to the change of the
lateral deflection at the rear (v,.,,). The contact-length-dependent coeffi
cients appearing in the expressions are:
A(a) = {-2e + -} /P(a)
B(a) - 2/P(a)
(9.5.134)
B*(a) = {(1 +

P(d) =
1-c

They are derived from solutions of the differential equation for the shape
of the string (9.5.35). For the parameters employed we refer to expressions
ANALYSIS OF TIRE PROPERTIES 855

(9.5.34). From the differences shown above (9.5.132-133) the following


differential equations for the unknowns νc1 and νc 2 are formed:

ds ds '-£ds
*<«) 1 + ^T a
(9.5.135)
dv« da \ v2 — vc2
l~ %-]--»*
ds ds a

B(a) 1 + -==- —+
ds
The two remaining unknowns ν1 and ν2, denoting the total lateral deflec
tions and consequently the distance between wheel plane and contact line
at the leading and trailing edges, are found by use of the following consid
erations. In the rolling process of a drifting tire with a continuously chang
ing contact length, in general, three intervals can be distinguished during
the contact phase of the loading cycle. In figure 9.5.44 a possible variation
of the contact length 2a has been shown in the road plane (x, y). The tire
touches the ground over a certain distance of travel. Immediately after the
first point touches the road, the contact line will grow in two directions.
This will continue until the second interval is reached, where growth of
contact takes place only at the front, and at the rear loss of contact occurs.
In the third interval, finally, loss of contact at both ends takes place until
the tire leaves the road. When the tire does not leave the road, an addi
tional interval II occurs before interval I is reached again and the cycle has
started anew. In less severe cases, intervals I and III may not occur. We
then have the relatively simple situation of continuous growth of contact
at the front and loss of contact at the rear.
The unknowns ν1 and ν2 and the ^-coordinate of the contact points in
the (x, v) plane are obtained as follows (cf. fig. 9.5.44).

da
Interval I: , > 0, Jc2 < 0 >l
ds
V| - Vcl, V2 = Vc2
• yri =sa- vfl, - ycl = sa- vcl

, da
Interval II: , > 0, Jc2 > 0 - 1 T~
ds (9.5.136)
vcl, v2 = sot + yc2
- yet = sa- vel

Interval III:
as
** V2
856 MECHANICS OF PNEUMATIC TIRES

FIGURE 9.5.44. The development of the contact line of the tire model which periodically loses
contact with the road.
Arrowi indicate the directions in which the positions of the leading and trailing contact points change.

By means of numerical integration, equations (9.5.135) have been solved


with the aid of the equations (9.5.136) and the expressions (9.5.134) for the
case formulated below in which the vertical tire deflection δ has been
given a periodic variation with amplitude δa greater than the static deflec
tion <*>„, so that periodic loss of contact between tire and road occurs. The
rated or static situation will be indicated with subscript o. The tire radius
is denoted by r. The vertical deflection and the contact length are gov
erned by the equations
S - 5,, - 8, sin UA (9.5.137)
S-IrS-S2. (9.5.138)
The numerical values of the system under consideration are as follows:

a0 = 0.25r, (80 = 0.03175r), 8. - O.lr, «, = f-


(9.5.139)
a = 3a0, e - 0.25, (a, = 0.75a.).
In figure 9.5.45 the calculated variation of contact length "La (oval curve),
the path of the contact points (lower curve AB) and the course of the
points of the string at the leading and trailing edge (AC and BC) are in
dicated. Two positions of the tire are shown, one in interval II and one in
interval III.
ANALYSIS OF TIRE PROPERTIES 857

It is necessary that the calculations in the numerical integration process


be carried out extremely accurately. A small error gives rise to a rapid
build up of deviation. In the case considered above, the integration time is
limited and accurate results can be obtained. In cases, however, where
continuous contact between tire and road exists, a very long integration
time is needed before a steady state situation has been attained, which is
due to the fact that the exact initial conditions of one loading cycle are not
known. For this sort of situation, the exact method described above will be
difficult to apply due to strong drift of v, 2 in particular.
For further investigation of the effect of a time varying load we will turn
to an approximate method based on the behavior of a string model with
out tread elements.

Simple string model


The treatment of this model is much simpler since the deflections at
both ends are independent of each other. Drift does not occur in the calcu
lation process. In figure 9.5.46a the deflected model has been shown in in
terval II. In the three intervals the following sets of equations apply.
Growth of contact at leading edge:

da
>0

dv{ _ _Vi_ da
ds s ds (9.5.140)
x^s + a,
- ycl = sa - v,.

FIGURE 9.5.45. The deflected lire model in two positions during the interval of contact with
the road.
858 MECHANICS OF PNEUMATIC TIRES

Growth of contact at trailing edge:

*•<"[%> l

a \ ds (9.5.141)

sa- v2.
Loss of contact at leading edge:

ds
(9.5.142)
+ yel.
Loss of contact at trailing edge:
da

(9.5.143)
v2 = sa + ycl.
Solutions of the above equations show considerable differences from the
results obtained using the more advanced model with tread elements. The
most important difference is the fact that with the simple model the lateral
force does not gradually drop to zero as the tire leaves the ground.
In order to get better agreement we introduce a relaxation length σ = σ*
(a) which is a function of the contact length 2a. We will take σ* equal to
the relaxation length of the more advanced model according to eq (9.5.47)
and figure 9.5. 18.
In figure 9.5.46b a comparison is made of the results for the three cases:
without tread elements, with tread elements (exact) and according to the
approximation with varying σ = σ* (a). The calculations are carried out
for the values (9.5.139). The approximate path shows good agreement with
the path of the contact points for the model with tread elements. When the
tread elements are omitted the path becomes wider and the lateral deflec
tions become greater.
The lateral force Ft. and the moment Mz which act on the tire can be de
termined with good approximation by the following simplified formulae.
In their derivations we have replaced the contact line by the straight line
connecting the beginning and end points of contact. For the model with
tread rubber elements we obtain:

(9.5.144)
a)C(a)} a.
ANALYSIS OF TIRE PROPERTIES 959

FIGURE 9.5.46a. The development of the contact line for the simple tire model without tread
rubber.

and for the model without tread elements:

(9.5.145)
M, = -2c,a (Yi <t + a (a + a)} a,.

tangtntt to path
exact without I elements
trailing path approx. with tread elements
leading edgt -path exact with tread elements

FIGURE 9.5Mb. Paths of contact points and variations of lateralforce Fy according to three
theories.
860 MECHANICS OF PNEUMATIC TIRES

Here we have introduced the quantities


v0 = V4 (v, + v2), a, - % (v, - v2)/a (9.5. 146)
and
a
C(a)-A(a)--cB*(a) 2 + - fB(a) (9.5.147)
a
in which A, B and B* are given by expressions (9.5.134).
For the sake of completeness the formulae for the moment have been
given. We shall restrict ourselves to discussion of the variation of the force.
For the three cases considered, the variation of the cornering force has
been shown in the same figure 9.5.46b. As expected, the approximate and
exact theories drop to zero at the point of lift off, whereas the force acting
on the simple string model remains finite, under the assumption of no slid
ing. The correspondence between exact and approximate solutions of the
path and of the force are satisfactory, and we will henceforth use the ap
proximate method exclusively for the investigation of the model with
tread elements.
For a series of amplitudes δa and path frequencies ωs the variation of the
cornering force, or rather of the cornering stiffness, has been calculated.
For illustration figure 9.5.47 gives the time histories of the cornering force
acting upon the model without tread rubber elements. Four cases are con
sidered. The upper figure shows the effect of a large amplitude, and it is
clearly seen that oscillations with the larger wavelength (λ) attain a higher
average side force. The effect of an amplitude equal to the static deflection
is presented in the second figure. Two possibilities are considered. When
contact is not lost, the tire lateral deflection does not need to be developed
anew each cycle from the undeformed state, which happens when the tire
leaves the ground. This causes the much lower level of the cornering force
shown in the latter case. The lower figure gives the force variation which
occurs at moderate values of the vertical amplitude.
Since during loss of contact negative vertical forces cannot be trans
mitted, the period of the total loading cycle must become greater in order
to keep the average vertical load unchanged. This change in period has
been taken into account in the calculation of the average side force or cor
nering stiffness CFa.
Figure 9.5.48 shows the final result of this investigation, viz the corner
ing stiffness averaged over one complete loading cycle as a function of ver
tical deflection amplitude δa and path frequency to, for both models with
and without tread elements. With the values σ = 3a0 and e = 0.25 the re
laxation length of the more advanced model becomes a* = 1.7a0. For the
model without tread rubber a relaxation length has been considered equal
to σ = 2a0.
The figure clearly illustrates the unfavorable effect of increasing the am
plitude and the path frequency. The curve at zero frequency is purely due
to the nonlinear variation of the cornering stiffness as a function of verti
cal load shown in figures 9.5.18 and 9.5.42. A pronounced difference be
tween the response of both types of models is the discontinuity which
takes place in the curves for the simple string model. This is in contrast to
the gradual variation in average cornering stiffness which occurs with the
more advanced model, and no doubt also with the real tire. The more ad-
ANALYSIS OF TIRE PROPERTIES 861

FIGURE 9.5 AT. Calculated lateralforce variation of the drifting simple string model (sliding
not considered) due to a periodically changing vertical load W.
The lower figure represents the cue of constant contact, the center figure the cues ofjust maintaining and losing contact
and the upper figure the cue of periodic loss of contact.

vanced model, furthermore, already shows a noticeable decrease in aver


age cornering stiffness before the tire periodically leaves the ground. The
very low level to which the average stiffness has been reduced, once loss of
contact occurs, is of the same order of magnitude for both tire models.
862 MECHANICS OF PNEUMATIC TIRES

SIMPLE STRING MODEL

V
amplitude of
deflection 8, -3 15 Bo

STRING MODEL WITH TREAD ELEMENTS

AVERAGE
CORNERING
STIFFNESS

FIGURE 9.5.48. Calculated variation of average cornering stiffness CFa with amplitude of tin
normal deflection Sa and path frequency.

Finite values of slip angle


Thus far we have only discussed the behavior at small slip angles, where
no sliding is assumed to occur. Larger slip angles will induce zones of slid
ing and cause a much more complicated situation. We have not attempted
to extend this investigation into the nonlinear range of the problem.
For the analysis of automobile motions we might use the experimentally
or theoretically determined average cornering stiffness function C^S., u.)
for the construction of the average tire characteristic /%(«) from the basic
characteristic F^(oeq) (cf. sec. 9.5.1) through the introduction of the equiv
alent slip angle
•a. (9.5.148)

A semi-empirical theory of the effect of a time varying load has been de


veloped by Metcalf [55]. In this theory large slip angles and the possibility
ANALYSIS OF TIRE PROPERTIES 863

of sliding have been included. Metcalf carried out a number of experi


ments with a 4.00-7 tire which was loaded against a rotating drum. During
each loading cycle, induced by a sinusoidally changing axle displacement,
two distinctly different phenomena were observed. First, as the normal
load increases from a minimum value, the lateral force lags behind the ap
plied vertical load. Secondly, after the vertical force has reached a maxi
mum and has begun to decrease, a point is reached where the lateral force
becomes equal to the instantaneous load on the tire multiplied by the fric
tion coefficient between tire and drum.
This behavior has been simulated with the aid of an analog computer. A
first-order filter has been employed for the generation of the phase lag.
The response to a step change of the input corresponds to the lateral force
response of the tire to a step change in vertical load (equal to the response
shown in fig. 9.5.34) approximated by:
F, = F^(l - «-"•). (9.5.149)
Although Metcalf recognized that the relaxation length a changes with
vertical tire deflection, in the computations a has been held constant for
the sake of simplicity. By means of a "least selector," the lateral force has
been limited to the friction coefficient multiplied by the normal load
In figures 9.5.49-54 Metcalfs computer and experimental results are
presented. In the simulation a number of constant values of a have been
tried. In figure 9.5.49 the dynamic loss, which together with the static loss
indicated in figure 9.5.42 forms the reduction in lateral force due to load
variations, has been plotted against reduced or path frequency u, for three
drum velocities. The test data from this figure support the hypothesis that
the phenomenon is distance rather than time dependent. In figure 9.5.50
the average lateral force fy has been plotted against reduced frequency u,
for two values of slip angle. The amplitude of the tire vertical deflection is
lower than the static deflection so that contact between road and tire is
maintained. In figure 9.5.51 results are plotted for the same tire but with
higher inflation pressure and an amplitude greater than the static deflec
tion, which results in a periodic loss of contact. Completely in accordance
with the theory (cf. also fig. 9.5.48), the reduction with increasing reduced
frequency w, becomes stronger for amplitudes in excess of the static tire
deflection.
In figures 9.5.52-54 the time histories are shown for this Utter configu
ration. These plots show agreement between test data and computed re
sults. The lateral resonance of the measuring axle accounts for the high
frequency oscillations present in the measured time histories.
The decrease of Fy with u, has been reasonably well predicted by the
analog computer results for σ = 18 cm. except at high values of slip angle
where tests show that a sharper decrease of Fy with u, occurs. This value
o=l8 cm. is approximately 20 percent larger than the value obtained ex
perimentally at rated constant deflection Wo and at small slip angles.
Interaction of vertical and horizontal motions
Hitherto the .slip angle of the tire has been considered as constant. In
reality, however, the mass of the automobile is not infinitely large. Due to
the variation of the lateral force being caused by a time-varying load, the
wheel plane will move laterally. This brings us to the problem of com
bined nonstationary tire behavior influenced by both the vertical and hori
zontal wheel motions.
864 MECHANICS OF PNEUMATIC TIRES

z
>. „
U. iW
400.7tyra
<
in • Pi.Wbar
i
M 0 °
0 A> load W0 .1340 N
OfcampK, ,127cm
o Stiir angltot.S*
1 O V . 17.S km/h
•S 5 Sp**d«, V .43 km/h
in XX) D V 7C km/h
U
o
E• 0 A
-^

c
/T

^
^
0
10 20
Reduced loading frequency. cus (rad/m)

FIGURE 9.5.49. Dynamic side force loss versus reduced loading frequency for three drum
velocities.
(Pip. 9.J.49-M ire taken from Metcalf [55])

In the literature, this problem has been touched at by Böhm [49], He


presents an approximate theory of the lateral motion of the mass m rigidly
connected to the king-pin of the wheel which shows a constant steer angle
if (cf. fig. 9.5.55). The mass is subjected to a constant lateral force K repre
senting, for instance, the centrifugal force. Böhm considers a periodically
changing vertical load and cornering stiffness, from which follows, accord
ing to eq (9.5.51), that for a constant lateral stiffness Ct the relaxation

400.7tyrt
. 1.4 bar

Reduced loading frequency, uMrad

FIGURE 9.5.50. Average side force versus reduced loading frequency-


ANALYSIS OF TIRE PROPERTIES 865

1000 4.00 - 7 «»rt


p. • 2.8 bar
Q miafurtd
I Av load W0
—computed
2 Av dtli J, « O.J em
•H" Die. ampl.8, = 1.3cm
o> Sltcr angle a > 5°
o
Spc.d V > 17.5 Km /h

10 40
Reduced loading frequency, wt(rad/m)

FIGURE 9.5.5 1. Average side force versus reduced loading frequency.

length a* also changes periodically. The following formulae are assumed


to hold:
W
(9.5.150)

__O Ten data


Analogue rtiultl,
Ailc displacement
to < 2.7 rad/« . OJ,= 0

V
E

u
JJa. 1

_. 0
x

-1

FIGURE 9.5.52. Time history, configuration offigure 9.5.51, low frequency.


866 MECHANICS OF PNEUMATIC TIRES

r —Q--TOI data
y1 Analogue rttultt (Tiltcm

li

FIGURE 9.5.53. Time history, configuration offigure 9.5.51, moderate frequency.

From these assumptions it may be noted that the average values are not
constant but increase with f !
The following simple differential equation for the nonsteady state varia
tion of the lateral deflection ν of the tire, valid for vanishing size of the

c
«l
s,

-500

0.13 016

Time (sec.)

FIGURE 9.5.54. Time history, configuration offigure 9.5.51, high frequency.


ANALYSIS OF TIRE PROPERTIES 867

FIGURE 9.5.55. The simple tire model usedfor the analysis of interaction between vertical and
lateral wheel motions.

contact area, has been applied:

J_ 1 + f sin at
V V, (9.5.151)

The equation of motion for the mass reads:


my Crv - K. (9.5.152)
Elimination of y yields:

Jsinwf) — v v=—• (9.5.153)


m m
This equation with periodically changing coefficients may be solved with
the aid of the perturbation method. We consider f small and put
v = v, + fv,(0 + (9.5.154)
After insertion of this power series in (9.5.153) another series arises which
must vanish identically in f, hence the coefficients of the successive powers
of f must vanish.
When we restrict ourselves to the second power of f the following form
is obtained for the lateral tire deflection:
v = \0 + £4, cos (at - <f>,) + l?A2 cos (2«f - <fc) + ••• (9.5.155)
For the factor of the first harmonic we obtain:
aV_
a?
tan $, (9.5.156)
-w2

It may be noted from these results that the maximum variations of lateral
deflection are obtained at a frequency of the load variation equal to the
natural frequency of the horizontal motion (ω2 = CJm). Insertion of this
868 MECHANICS OF PNEUMATIC TIRES

solution in equation (9.5. IS 1) yields for the velocity of the mass in the lat
eral direction (assumed equal to zero for f = 0):
y sin2 ut • sin £, + f2^ sin (2w/ - i^) + —
-y = SB, sin (ut - $,) + fj —At
(9.5.157)
The average of this expression does not vanish. As A1 is negative (9.5.156)
the average cornering stiffness apparently increases with f (this is in con
tradiction with the result of Böhm, presumably due to sign error). The
maximum average lateral velocity

-*r». (9.5.158)

exactly equals the decrease in side slip velocity resulting from the rise in
average cornering stiffness due to the assumed variations (7.5.150). A bet
ter assumption keeping the average load constant no doubt will lead to
different results.

References
[1] Joy, T. J. P., and Hartley, D. C, Tire characteristics as applicable to vehicle stability
problems, Proc. IME, Auto. Div. (1953-1954), p. 113.
[2] Gausz, F., and Wolff, H.. Ueber die Seitenführungskraft von Personen-wagen-Reifen,
Deutsche Kraftf. forsch. und Sir verk. techn., Heft 133 (1959).
[3] Fonda, A. G., Tire tests and interpreation of experimental data, Papers on research in
automobile stability and control and in tire performance, Aut. Div. of IME, No. 7
(1956/57).
[4] Freudenstein, G., Luftreifen bei Schräg- und Kurvenlauf, Deutsche Kraftf forsch und
Str. verk. techn., Heft 152 (1961).
[5] Nordeen, D. L., and Cortese, A. D., Force and moment characteristics of rolling tires,
SAE Paper No. 713A (June 1963); SAE Trans., 325 (1964).
[6] Henker, E., Dynamische Kennlinien von PKW-Reifen, Wissenschaftlich-Technische
Veröffentlichungen aus dem Automobilbau (IFA-DDR) Heft 3 (1968).
[7] Fonda, A. G., and Radt, H., Summary of Cornell Aeronautical Laboratory tire test
data, CAL Report YD-1059-F-2 (1958).
[8] Smiley, R. F., and Home, W. B., Mechanical properties of pneumatic tires with special
reference to modem aircraft tires, NACA (NASA) Tech. Note 4110 (1958).
[9] Haedekel, R., The mechanical characteristics of pneumatic tires. S & T Memo No. 10/
52 (British Ministry of Supply, TPA 3/TIB, 1952).
[10] Smiley, R. F., Correlation, evaluation, and extension of linearized theories for tire mo
tion and wheel shimmy, NACA (NASA) Tech. Note 3632 (June 1956).
[11] Frank, F., Grundlagen zur Berechnung der Seitenführungskennlinien von Reifen,
Kaut. Gurnmi 8 (18), 515 (1965).
Frank, F., Theorie des Reifenschraglaufs, Diss. T. H. Darmstadt (1965).
Fiala, E., Seitenkräfte am rollenden Luftreifen, VDI-Zeitschrift 96, 973 (1964).
Ellis, J. R., Frank, F., and Hinton. B. J., The experimental determination of tire model
parameters, A.S.A.E. Report No. 2 (Sept. 1966).
[15] Böhm, F., Der Rollvorgang des Automobil-Rades, ZAMM 43, T56-T60 (1963).
[16] Borgmann, W., Theoretische and Experimentelle Untersuchungen an Luftreifen bei
Schräglauf, Diss, Braunschweig (1963).
[17] Bergman, W., Theoretical prediction of the effect of traction on cornering force, SAE
Trans., 614 (1965).
[18] Pacejka, H. B., Study of the lateral behaviour of an automobile moving upon a flat level
road, Cornell Aeronautical Laboratory Report YC-857-F-23 (1958).
[19] Pacejka, H. B., The wheel shimmy phenomenon, Diss., Tech. University of Delft
(1966).
[20] Rotta, J., Zur Statik des Luftreifens, Ing. Archiv, 129 (1949).
ANALYSIS OF TIRE PROPERTIES 869

[21] Kalker, J. J., On the rolling contact of two elastic bodies in the presence of dry friction,
Diss., Delft (1967).
[22] Nayak, P. R , and Paul, I. L., A new theory of rolling contact, Engrg. Proj. Lab., Dept.
of Mech. Engrg., M.I.T. (1968).
[23] Savkoor, A. R., The lateral flexibility of pneumatic tire and its application to the lateral
rolling contact problem, SAE Paper No. 700378 (1970); FISITA Int. Auto. Safety
Conf. Compendium, p. 367, New York (1970).
[24] Savkoor, A. R., The relation of the adhesional friction of rubber to the friction between
tire and ground, Paper B-12, FISITA, Mflnchen (1966).
[25] Schlippe, B. von, and Dietrich, R., Das Fattern eines bepneuten Rades, Ber. Lilienthal
Ges. 140, p. 35 (1941).
[26] Schlippe, B. von, and Dietrich, R., Shimmying of a pneumatic wheel (Translation of
[25]), NACA (NASA) TM 1365, p. 125 (1954).
[27] Holmes, K. E., and Stone, R. D., Tire forces as functions of cornering and braking slip
on wet road surfaces, Paper 6 (IME Auto. Div. Symp., Handling of Vehicles under
Emergency Conditions, Univ. of Technology, Loughborough, England, Jan. 8, 1969),
Proc. IME 1968-69 183 (Pt 3H), 35 (1969).
[28] Gengenbach, W. (Private Communication).
29] Riekert, P., und Schunch, T. E., Zur Fahrmechanik des Gummi-bereiften Kraftfahr-
zeugs, Ing. Archiv 11, 210 (1940).
[30] Whitcomb, D. W., and Milliken, W. F., Design implications of a general theory of auto
mobile stability and control. Papers on research in automobile stability, Proc. Auto.
Div. of IME, No. 7 (1956-57), with many references.
[31] Segel, L., Theoretical prediction and experimental substantiation of the response of the
automobile to steering control, Proc. Auto. Div. of IME, No. 7 (1956-57).
[32] Mitschke, M., Fahrtrichtungshaltung, Analyse der Theorien, A.T.Z. 70(5), 157 (1968),
with many references.
[33] Fiala, E., Zur Fahrdynamik des Strassenfahrzeuges unter Berücksichtigung der Len-
kungselastizität, A.T.Z., 71 (1960).
[34] Pevsner, Ja. M., Theory of the stability of automobile motions (in Russian), Masjgiz,
Leningrad (1947).
[35] Hoffmann, E. R., Note on vehicle stability. Austral. Road Research 1, (1964) III, p. 15.
[36] Antonov, D. A., Stability calculation of automobiles (in Russian), Avtomobiljnaja
promisjlennost. No. 9 (1963).
[37] Apetaur, M., Beurteilung der Fahreigenschaften von Fahrzeugen nach der Ergebnissen
der Prtifungen durch gleich massige. Fahrt auf einer Kreisbahn, Conference on Test
ing of Automobiles, Prague (Oct. 1965).
[38] Radt, H. S., and Pacejka, H. B., Analysis of the steady-state turning behavior of an au
tomobile, Proc. Symp. Control of Vehicles 1963, p. 66 (IME, London).
[39] Böhm, F., Ueber den Fahrzustand des Kraftwagens auf einer ebenen Kreisbahn ohne
Ueberhöhung, Ing. Archiv. 32, 112 (1963).
[40] Kiiter, W. T., and Pacejka, H. B., On the skidding of vehicles due to locked wheels. Pa
per 1 (IME Auto. Div. Symp., Handling of Vehicles under Emergency Conditions,
Univ. of Technology, Loughborough, England, Jan. 8, 1969), Proc. IME 1968-69, 183
(Pt 3H), 3 (1969).
[41] Chiesa, A., Rinonapoli, L., and Bergoni, P. I. R., A new loose inverse procedure for
matching tires and car using a mathematical model, Paper 4 (IME Auto. Div. Symp.,
Handling of Vehicles under Emergency Conditions, Univ. of Technology,
Loughborough, England, Jan. 8, 1969), Proc. IME 1968-69, 183 (Pt 3H), 93 (1969).
[42] Harris, A. J., and Riley, B. S., Vehicle behavior in combined cornering and braking. Pa
per 5 (IME Auto. Div. Symp., Handling of Vehicles under Emergency Conditions,
Univ. of Technology, Loughborough, England, Jan. 8, 1969), Proc. IME 1968-69 183
(Pt 3H), 19 (1969).
[43] Hllis. J. R., Understeer and oversteer, Auto. Engr. (London) (May 1963).
[44] Kantrowitz, A., Stability of castering wheels for aircraft landing gears, NACA (NASA)
Report 686 (1937).
[45] Greidanus, J. H., Besturing en stabiliteit van het neuswielonderstel, NLL Report V
1038, Amsterdam (1942).
[46] Saito, Y., A study of the dynamic steering properties of tires, FISITA London (1962),
pp. 101, 246, 282.
[47] Fromm, H., Kurzer Bericht ueber die Geschichte der Theorie des Radflatterns, Ber. Li
lienthal Ges. 140, 53 (1941); or NACA (NASA) TM 1365, p. 181 (1954).
[48] Bourcier de Carbon, C., Etude theorique due shimmy des roues d'avion, Off. Nat.
d'Etude et de Rech. Aer. 7 (1948); or NACA (NASA) TM 1337 (1952).
[49] Bflhm, F., Reifenmechanik und fahreigenschaften des automobils, Paper B-3, FISITA
München (1966).
870 MECHANICS OF PNEUMATIC TIRES
[SO] Schlippe, B. von, and Dietrich, R , Zur Mechanik des Luftreifens, Zentral für Wiss.
Ber., Berlin (1942).
[51] Schlippe, B. von, and Dietrich, R., Das Flattern eines mil Luftreifen versehenes Rades,
Jahrbuch Deutsche Luftfahrtforschung (1943).
[52] Segel, I.., Force and moment response of pneumatic tires to lateral motion inputs,
Trans. AS ME. J. Engr. for Ind. 88B(1) (1966).
[53] Kurz, H., Seitenführungskraft des Kraftwagenrades bei wechselnder Radlast, A.T.Z..
60(5) (1958).
[54] Endres, W., Versuche ilber das Verhalten des Autorades in der Kurve, VDI-Zeitschrift
106(4) (1964).
[55] Metcalf, W. H., Effect of a time-varying load on side force generated by a tire operating
at constant slip angle, SAE Paper No. 713C (1963).
[56] Pacejka, H. B., Discussion, Proc. Symp. Control of Vehicles during Braking and Cor
nering 1963, p. 1 16 (IME, London).
[57] Ginn, J. L., Miller, R. F., Marlow, R. L., and Heimovics, J. F., The B. F. Goodrich Tire
Dynamics Machine, SAE preprint 490 B (March 1962).
[58] Pacejka, H. B., Principles of plane motions of automobiles, Proc. IUTAM Symp. on
Dynamics of Vehicles, ed. H. B. Pacejka, Delft, Aug. 1975, p. 33.
[59] Sorgatz, U., Ein theoretisches Fahrzeug Modell zur Abbildung der Fahrdynamik bis in
den Grenzbereich, Doctoral thesis, Aachen Univ. of Technology, 1973.
[60] Pacejka, H. B. and Fancher, P. S , Hybrid simulation of shear force development of a
tire experiencing longitudinal and lateral slip, Proc. XIV Int. Auto. Tech. Congress,
FISITA, London, June 1972, p. 3/78.
[61] Pacejka, H. B., Simplified analysis of steady-state turning behaviour of motor vehicles.
Vehicle System Dynamics 2 (1973) p. 161, 173, 185.
[62] Rogers, L. C. and Brewer, H. K. Synthesis of tire equations for use in shimmy and other
dynamic studies, J. of Aircraft, Vol. 8, No. 9, Sep. 1971, p. 689.
[63] Rogers, L. C., Theoretical tire equations for shimmy and other dynamic studies, J. of
Aircraft, Vol. 9, No. 8, Aug. 1972, p. 585.
[64] Pacejka, H. B., Analysis of the dynamic response of a rolling string-type tire model to
lateral wheel-plane vibrations, Veh. Sys. Dyn. 1 (1972), p. 37.
[65] Pacejka, H. B., Approximate dynamic shimmy response of pneumatic tires, Veh. Sys.
Dyn. 2(1973), p. 49.
[66] Segel, L. and Wilson, R., Requirements for describing the mechanics of tires used on
single-track vehicles, Proc. IUTAM Symp. on Dynamics of Vehicles, ed. H. B. Pa
cejka, Delfl, Aug. 1975, p. 173.
[67] Ho, F. H. and Hall, M. F., An experimental study of the pure-yaw frequency responses
of the 18 X 5.5 type VII aircraft tires, AFFDL-TR-73-79, Air Force Flight Dyn. Lab.,
Wright-Patterson Air Force Base, Ohio 1973, 138 p.
[68] Sekula, . J., Hall, G. L., Potts, G. R. and Conant, F. S., Dynamic indoor tire testing and
Fourier transform analysis. Tire Sc. and Tech., Vol. 4, No. 2 (May 1976) p. 66.
[69] Weber, R., Der Kraftschluss von Fahrzeugreifen und Gummiproben auf vereister
Oberfläche, Doctoral Thesis, Karlsruhe University, 1970.
[70] Willumeit, H. P., Theoretische Untersuchungen an einem Modell des Luftreifens . . . ,
Doctoral Thesis, Tech. University Berlin, 1969.
[71] Pacejka, H. B., Some recent investigations into dynamics and frictional behavior of
pneumatic tires. Proc. GM. Symp. Physics of Tire Traction, eds. D. F. Hays and A. L.
Browne, Plenum Press, New York 1974, p. 257.
[72] Koch, B., Computer simulation of steady-state and transient tire traction performance.
Proc. IUTAM Symp. on Dynamics of Vehicles, ed. H. B. Pacejka, Delft, Aug. 1975,
p. 197.
[73] Krempel, G., Untersuchungen an Kraftfahrzeugreifen, A. T. Z., Vol. 69, Nos. 1 and 8,
1967, p. 1-8, 262-268. (cf. also Dissertation Karlsruhe University, 1965).
[74] Holmes, K. F. and Stone, R. D., Tire forces as functions of cornering and braking slip
on wet road surfaces, Proc. I.M.E., Vol. 183, Part 3H, 1969, p. 35.
[75] Fancher, P. S., Segel, L., MacAdam, C. and Pacejka, H. B., Tire traction grading proce
dures as derived from the maneuvering characteristics of a tire-vehicle-system, VoL I
& II, final report, NBS Contract No. 1-35715, H.S.R.I., University of Michigan, Ann
Arbor, June 1972.
[76] Segel, L., Tire traction on dry uncontaminated surfaces, Proc. G. M. Symp. Physics of
Tire Traction, eds. D. F. Hays and A. L. Browne, Plenum Press, New York 1974, p. 65.
[77] Gengenbach, W , Experimentelle Untersuchung von Reifen auf nasser Fahrbahn, ATZ
70 (1968) 1, p. 83; 2, p. 288; 3., p. 310.
Chapter 10
PHYSICAL PROPERTIES OF
RUBBER COMPOUNDS
J. R. Beatty1

PIP
10. 1 Introduction 872
10.2 Characterization of Tire Compounds 875
10.3 Rubber As An Engineering Material 875
10.4 Relationships Between Properties Measured in
Tension, Compression and Shear 877
10.5 Discussion of Properties and Their Significance 880

1 The B.F. Goodrich Research and Development Center 9921 Brecksville Road, Brecksville. Ohio 44141.

871
872 MECHANICS OF PNEUMATIC TIRES

10.1 Introduction

For many years pneumatic tires were constructed by somewhat Edi-


sonian approaches. The cord and rubber materials were selected from ex
perience and the construction details of the tire were arrived at largely by
cut-and-try methods. But in the past two decades advances have been
made in both the physical understanding and the mathematical modeling
of tire performance. Several reasons exist for this change. One is the emer
gence of composite materials as a branch of mechanics, partly due to im
petus from aerospace applications, which has resulted in a better under
standing of cord-rubber composites used as structural materials in tires.
Another is the rapid development in instrumentation and measurement
technology, which has given the tire engineer far more data concerning his
product than he has ever had before. Another is the opening up of the rel
atively new areas of polymer friction and wear. A better understanding of
these mechanisms has in turn led to increased traction and Life of pneu
matic tires. Finally the emergence of high speed digital computation has
for the first time made possible the accurate computation of stress, defor
mation and temperature in a tire, as well as aided immensely in data re
duction and analysis.
Bias and radial tires require different rubber compounds in order to per
form properly in service. A bias tire requires as a minimum about half a
dozen rubber compounds, while a radial truck tire may have as many as a
dozen. Each compound has different physical properties depending on its
use as a tread, sidewall, ply stock, overhead or belt stock, bead profile,
filler ply, undertread, liner, bead filler, chafer, gum strip or decorative
sidewall. Each rubber compound has its function, and for the most part
these functions are reasonably obvious. Figure 10. 1a illustrates where the
compounds are usually located in a bias passenger tire, while Figure 10.1b
illustrates this for a truck tire. Figure 10.2a shows the location of com
pounds in a radial passenger tire, and Figure 10.2b in a radial truck tire.
There is no general agreement within the tire industry as to how many dif
ferent rubber compounds to use, or what their composition should be.
However, in surveying the literature on tires from various sources, it is ob
vious that many manufacturers have learned by either physical measure
ments or by experiment that certain values of properties are required in
specific components of the tire.
The purpose of this chapter is to present the mechanical and physical
properties of rubber compounds most important to tire performance. The
values listed do not represent the properties of any one tire or manufac
turer, but are instead typical values as might be found by averaging a
number of tires from different manufacturers. This data should be useful
for mathematical modeling of tire performance, and for an understanding
of the range of properties in a modern tire.
The cost of the materials, including cord, is approximately two thirds of
the cost of the finished tire, so the rubber compounder must constantly
strive for maximum cost effectiveness. However, the physical properties
are the criteria for the adequacy of a compound, and some general obser
vations on the types of rubber and carbon black used for the different
components may be of interest here.
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 873

TREAO

CHAFED
BEAD

BIAS-PLY PASSENGER TIRE

FIGURE 10. la Location of the Components in the Cross-Section of Bias Ply Passenger Tire

TREAO

UNOER-TREAO

SIDEwan
com PLIES (6)
(PLY STOCK]

BIAS-PLY HEAVY DUTY TIRE

FIGURE lO.lb Location of the Components in the Cross-Section of Bias Ply Truck Tire
874 MECHANICS OF PNEUMATIC TIRES

TREAD

STEEL BELTS (Z)


(PLY STOCK)

SIDEWALL

RADIAL PLIES (21

BEAD

STEEL BELTED RADIAL PASSENGER TIRE

FIGURE 10.2a Location of the Components in the Cross-Section of Steel Belted


Passenger Tire

s&gr
SKttMU

STEEL RADIAL HEAVY DUTY TIRE

FIGURE 10.2b Location of the Components in the Cross-Section of Steel Belted Radial
Truck Tire
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 875

10.2 Characterization of Tire Compounds


The proportion of natural rubber or its synthetic counterpart, cis-poly-
isoprene, is greater in the radial than the bias tire due to the unique prop
erties of natural rubber. Carbon blacks giving high modulus are normally
used with the primarily natural rubber compounds in radial tires.
The treads of passenger tires, whether radial or bias, are synthetic rub
bers, usually SBR:BR (styrene butadiene rubber: butadiene rubber)
blends. The proportions of rubber and the type of carbon black are se
lected to give the best balance of performance and durability to the tread.
In truck tires, depending on the service, either tread compounds similar to
passenger treads or natural rubber with minor proportions of cis-poly-
butadiene are used.
Due to its higher circumferential compliance, the sidewall of a radial
tire is subjected to different strain cycles than a bias tire, so compounds for
sidewalls of radial tires are formulated for greater fatigue resistance than
those for bias tires. All sidewall compounds are chosen to resist fatigue,
oxidative degradation, and ozone attack. To obtain this balance of proper
ties, blends of natural rubber, SBR and cis-polybutadiene are combined
with antioxidants, antiozonants, waxes and carbon blacks; the latter are
selected which give low modulus and hysteresis. White sidewalls require
non-staining properties as well as protection from environmental aging, so
rubbers resistant to oxidation and ozone attack are included in the blend.
Halogenated butyl and ethylene propylene diene rubbers are examples of
rubbers inherently resistant to oxygen and ozone without the addition of
staining antioxidants and antiozonants. Unfortunately the better anti
oxidants and antiozonants are staining materials.
Bead construction normally utilizes synthetic compounds in bias and
truck passenger tires, but more natural rubber is often used in radial tires
to achieve the desired properties.
Carcass coat stocks in bias passenger tires are blends of natural rubber
and various synthetics. In many instances, some reclaim rubber is used
since it does not detract from performance, and aids processing and re
duces cost. Carbon blacks are of the low modulus types. Bias truck tire
carcass stocks are quite similar to those of bias passenger tires except for
improved tensile strength and fatigue properties, which are normally ob
tained by increasing the proportion of natural rubber.
Most manufacturers utilize the low air permeability of butyl rubber
and/or butyl reclaim for the inner liner of tubeless tires. Inner tubes are
made of butyl or halogenated butyl, depending on the severity of service,
for the same reason. Carbon blacks used in these components are those
imparting low modulus and good processing characteristics.
10.3 Rubber as an Engineering Material
Rubber has several properties diffferent from the usual structural mate
rials. The first is its elastic deformability, which is orders of magnitude
greater than for metals or even wood. The second is that the tensile
strength and modulus of rubber are quite low, and the stress-strain curve
is nonlinear for large strains. A tensile stress-strain curve for a typical tire
rubber compound is shown in Figure 10.3. The tangent modulus, or slope,
of the stress-strain curve varies considerably. A third characteristic of rub
ber is the absence of a yield point in the stress-strain curve, unlike most
876 MECHANICS OF PNEUMATIC TIRES

conventional construction materials. A fourth is its ability to recover from


large deformation upon removal of the stress. Rubber has a remarkably
high energy-storage capacity and good fatigue properties, making it
ideally suited for use as springs, tires, and vibration isolating mountings of
all types.
Table 10.1 is quoted from Allen [1] in order to illustrate the energy-ab
sorbing capacity of rubber compounds compared to various other materi
als of construction. The extremely high value of energy absorbing capacity
for vulcanized natural rubber compared with the other materials is due to
its extremely high elasticity. Rubber is sometimes thought of as perfectly
elastic; however, it also has a viscous component that manifests itself as
creep or stress-relaxation, as well as in internal heat generation during cy
clic deformation. This heat generation due to hysteresis is discussed in
Chapter 3 and is often expressed quantitatively by tan δ' in Tables 10.3-
10.6.
The physical properties of rubber vary more rapidly with temperature
than do those of most conventional structural materials. They also vary
with frequency, as shown by Williams, Landel and Ferry, [2], with in
creasing frequency affecting the physical properties in the same manner as
decreasing temperature. Gehman discusses these changes in Chapter 1 of
this book, and the reader is referred to that chapter.
It is fortunate that the changes in modulus are small over the temper
ature range in which the rubber compounds in a tire operate, and further
that hysteresis decreases with increasing temperature. For the typical tire
the figures quoted will not change appreciably from the static tire condi
tion to the rolling tire state. One property of the tire, rolling resistance, de
creases somewhat during initial tire warm-up since tan δ decreases with
temperature, as does modulus to a lesser extent.
It is characteristic of rubber compounds that their physical properties
are not constant. Specifically, modulus and hysteresis are a function of the
age and strain history of the specimen. Modern compounds are not af
fected markedly by aging in the time frame used in tire mechanics, but the
strain history is important, and to further complicate the picture the effects

0 STRAIN——

FIGURE 10.3 Stress-Strain Curve in Tension for Typical Rubber Compound


PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 877

TABLE 10.1 Energy-Absorbing Capacity of Materials


Energy-absorbing Capacity
Joules/gram material Ft-lbs/lb material
Cast iron 0.0011 0.37
Soft steel 0.0092 3.1
Phosphor bronze 0.012 4.1
Rolled aluminum 0.023 7.6
Spring steel 0.29 95
Hickory wood 0.37 122
Vulcanized natural rubber 44. 14,700

are recoverable to varying degrees dependent on time and temperature.


Mullins [3] and Payne [4] have discussed the effects of strain history. They
found that all rubber compounds soften with repeated cycles of strain.
This effect is much more apparent in rubber reinforced with carbon black,
and the more reinforcing black* the greater the effect. This is one of the
reasons that tires are broken-in prior to measurements being made on
them. Break-in also allows the highly redundant cord and rubber com
ponents in a tire to equilibrate somewhat, which markedly improves du
rability. For example, a tire broken in under moderate conditions will not
fail nearly so quickly under severe test conditions as one not broken in.
All of the dynamic data given in this chapter are equilibrated proper
ties, except for ultimate tensile strength and static Young's modulus.
These characteristics of rubber require specific definition of the method of
measurement.
Poisson's ratio for rubber is very close to 0.5. This ratio varies slightly at
large deformations, which are seldom encountered in service, and is some
what less than 0.5 for "hard rubber" or ebonite. This value compares with
a Poisson's ratio of approximately 0.3 for metals. A Poisson's ratio of 0.5
means that rubber is incompressible, a fact which must be considered in
design for compression applications, since space for bulging of the sides of
the part is necessary when a rubber object is compressed.

10.4 Relationships Between Properties Measured hi


Tension, Compression and Shear
Tables 10.3-10.6 give common rubber properties. Durometer hardness
is measured by ASTM Method D-2240-73 [3] and has been shown by
Scott [5] Soden [6] and Gent [7] to be a measure of the modulus of the
rubber compound at low deformation. This test is extremely operator-sen
sitive and is useful only for quality control and/or approximate in
dications of stiffness. Tensile strength is computed using the original cross-
sectional area of the specimen.
Young's modulus is usually defined as the tangent, or slope, of the lin
ear portion of the stress-strain curve in the form
change in stress
change in strain
* The term "reinforcing black" is used to denote a carbon black with small particle size which markedly increases modu
lus and wear resistance. Larger particle non-reinforcing blacks are used primarily as cost reducing fillers or as processing
878 MECHANICS OF PNEUMATIC TIRES

and is a widely used measure for all materials. It is not widely used by
rubber technologists, who normally quote the "100%", 200% - 500%
modulus, which is of no physical significance since it is the stress at the
given elongation, and is similar to a secant modulus. These moduli do not
describe the shape of the stress-strain curve, so are of doubtful value in en
gineering applications. Young's moduli tabulated here were taken at small
strains, of the same order as strains in a tire, and were obtained from the
linear region of stress-strain curves, as shown in Fig. 10.4. On the other
hand, shear stress-strain curves are quite linear, and unless very precise
values are required, the magnitude of the strain does not affect the modu
lus significantly. Figure 10.5 illustrates a typical stress strain curve for
shear.
The relationship between Young's modulus "E" and shear modulus
"G" can be shown to be
E = 2G (1 + n) for small strains, where
Poisson's ratio /i = 0.5, so that

This relationship has been verified experimentally for a wide range of


compounds with varying moduli, and has been used in deducing values
for shear modulus of some of the tire compounds where direct measure
ments were not available in the literature.
Dynamic properties of rubber are important in engineering applications
since this is the manner in which the rubber is most widely used. Gehman
in Chapter 1 discusses dynamic properties and their importance in great
detail, and the reader is referred to his discussion of such properties and
tests to measure them. The dynamic modulus is always greater than the
static modulus for the reasons explained by Gehman. A more complete
discussion of dynamic properties is also given in ref. [9]. For this dis
cussion it is sufficient to say that the ratio of dynamic to static modulus
may vary from slightly over 1 to 2 or more, depending primarily on the
hysteresis of the rubber compound. Hysteresis is a function of polymer,

-COMPRESSION EXTENSION •

STRESS-STRAIN of RUBBER in EXTENSION and COMPRESSION

FIGURE 10.4 Stress-Strain Curve in Tension- Compression for Typical Rubber Compound
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 879

STRESS-STRAIN of RUBBER in SHEAR

FIGURE 10.5 Stress-Strain Curve in Shear for Typical Rubber Compound

black loading, state of cure, test temperature and other factors. These have
been discussed in detail by Studebaker and Beatty [10].
The dynamic properties tabulated here were obtained according to
ASTM method D2231-71 [11] at 10 CPS, 10% static strain, and ±5% dy
namic strain. Reference to Wood [12] for some of the values is acknowl
edged.
Thermal conductivity is tabulated, since it is a basic property of rubber
compounds which is important both in their manufacture and use. Heat
transfer is important in vulcanization of a tire and it will also influence the
performance characteristics in the vulcanized state. The thermal con
ductivity of unvulcanized and vulcanized rubber is essentially the same.
Thermal conductivity is important to performance due to heat generation
of the rubber in cyclic deformation as a tire rotates. Rubber has low ther
mal conductivity, so heat generated in the interior of a tire is not easily
conducted to the outside surface where cooling can occur. If heat is not
adequately conducted from the thicker sections, early thermal degradation
failures can occur. Furthermore, thermal conductivity is a function of tem
perature, as shown in Figure 10.6 from work done at RAPRA [14]. In this
figure, the upper two curves are typical of the majority of tire compounds.
It is unfortunate that as temperature increases thermal conductivity de
creases for these compounds. Reference [11] shows that the presence of
carbon black has an equalizing effect on thermal conductivity, as shown in
Table 10.2.

TABLE 10.2 Thermal Conductivity


(W/M • °K)

Rubber Gum 50 phr black


NR 0.153 0.280
SBR 0.190-0.250 0.300
BR 0.210 0.320
IIR(BUTYL) 0.130 0.230

Measurements reported in Tables 10.3-10.6 are for the 30°C range, so at


higher temperatures encountered in tire service conductivity becomes
somewhat less.
880 MECHANICS OF PNEUMATIC TIRES

_o.w-

tt
£025- -60phr ISAF BLACK

.020-

015-

50 100 ISO 200


TEMPERATURE, "C

FIGURE 10.6 Thermal Conductivity vs. Temperature for Various Compounds

10.5 Discussion of Properties and Their Significance


In the introduction it was stated that different compounds were devel
oped so that specific parts of a tire would have the required physical prop
erties. It is almost if not totally impossible to optimize all desirable proper
ties of a given compound even if cost were not a factor. If one property is
optimized, others must be neglected to some extent. Every rubber com
pound is the best compromise with the important property or properties
optimized.
For a tread compound, abrasion resistance, wet and dry skid resistance,
crack growth resistance and low hysteresis are desirable in approximately
that order.
The undertread, if used, is formulated to have low hysteresis and to
bond well to the cord-rubber composite making up the carcass or body of
the tire.
The sidewall, whether it is white or black, is formulated for good fatigue
life, resistance to oxygen and ozone attack, and to have good molding
properties, with hysteresis and abrasion resistance secondary consid
erations.
The carcass coat stock for use in the textile-rubber composite should
have good flow properties, good adhesion, low hysteresis, and good fa
tigue. The carcass coat stock for belts with steel wire is usually of higher
modulus than that intended for use with textile cords. This is to give better
adhesion to the high modulus brass plated steel wire. The high modulus
rubber compounds also contribute to the desired stiffness of the rubber-
wire cord composite in the steel belt. High tensile and shear strength and
good fatigue resistance are important properties.
The bead filler, and rubber compounds in the bead area, emphasize
high modulus and good adhesion to the exclusion of many other physical
properties which are not particularly important in the overall performance
of the tire. The innerliner of a tubeless tire requires good flex resistance,
particularly in radial tires. It must mold with no imperfections, and is usu
ally formulated to have low air permeability.
Tables 10.3-10.6 summarize the physical properties of the critical com
ponents of the various types of tires thought to be important in tire per
formance.
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 881

dddddddd

— — OO—"<
o
3

I — ooo — — — —

rjooooooor-
^•O — ^"ow^O^O

•oP

JS
Dynamic Modulus
•*<
E*
I 1
IVi
•f* w> O O O O i
^ m f^ —' ^ 1^ '

w^O^^W^^A
S oa

2 _u

.11

* IN —

**J
o

n .s.s
I!
I s
!?

a
§
H
ili
</5? fl2 oS 06 J
882 MECHANICS OF PNEUMATIC TIRES

O O O O O

i2§~
o~
.*<£ (N O ^ oo oo
IN r4 —'• ri <N
o
3 O O Q
rj oo O
ffl
1
£
I
Q r*% O O
O n^ vi m
fS — (N (N

o o o o d

u
^
«
11 O O O O
*r> irt o O
^ OO 55 (N

.•» a
U 00 —
— — 00 fN <S
^t ^' ri wS wS
~ C 3
« 3 TJ

*j|

•c
as
888SS

(N OO OO OO OO
^ o o o o I
<S <N (N <S (N

I
o
u
II:
s
9 o
O O
vO 00 .s s

!•§
!8

fi
P
lllii
H DCACO M
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 883

oooooo o

w"t »/i
10 CN v^
(N in
O trt
fS »A
O </•*
»/%

A oo r^ i^ w^ oo »n ^t
d d d d o d —

r- o o r~ o i
— o 5 >o — '

a,
•s
a HW d ddd d°°

|J

1J 8

5||
§88|g§«
I*Jg
aj 00

i c*"j OO OO OO OO O"1 OO
o

I 11 X
a
!
11
1?
el 31 .as
1=
II
a a

11

3& IIS*
•O
IjllJi
U w 4* ,9 fc-
H rt •§ o •- w « ^
illl^l s
HDBPE «•
884 MECHANICS OF PNEUMATIC TIRES

«:>• <N <N <N «S <S


O O O O O

VO « TT d in

o
3 r^ n o m c*»
ro fl O 00 m
<N (N <N — m

O — 00 Ov t
_ _ = .- -

I
o so
v* r- -"
o «*^
m Oo

r- >Q oo oo g
OO OOO

£ OO OO -" OO sO
^ ^ ^-' ro K
11
88838

•s-M
*il 383

O O O O
!»gg
09

2
o2o 52
o

<n M M M 2

c« O "-

II
Q
* .5.5
'H
13
II
1!
Co
Tire
11
H 3 c« E i
PHYSICAL PROPERTIES OF RUBBER COMPOUNDS 885

References
[1] Allen, P. W., Rubber Developments, 28, #1, 2, 1975.
2] Williams, Landel and Ferry, J. Am. Chem. Soc. 77, 3701, 1955.
ill (a) L. Mullins, Rub. Chem. and Tech., 21, 281, 1948.
(b) A. R. Payne, R. E. Whittaker, Rub. Chem. and Tech., 44, 440, 1971.
[4] ASTM D-2240-75, Annual Book of Standards, part 37, 608, 1976.
[5] J. R. Scott, Trans. Inst. Rub. Ind., 11, 224, 1935.
J. R. Scott, J. Rubber Res., 12, 117, 1943.
J. R. Scott, Ibid, 17, 145, 1948.
J. R. Scott, Ibid, 18, 12, 1949.
[6] A. L. Soden, A Practical Manual of Rubber Hardness Testing, Maclaren & Sons, Ltd.
Stafford House, Norfolk street, London, W.C. 2, England.
A. N. Gent, Rub. Chem. & Tech., 31, 896, 1958.
J. R. Beatty and M. L. Studebaker, Elastomerics, in press.
S. D. Gehman, Rub. Chem. and Tech., 30, 1202, 1957.
M. L. Studebaker and J. R. Beatty, Rub. Chem. and Tech., 47, 803, 1974.
ASTM D2231-71, Annual Book of Standards, part 37, 602, 1976.
L. A. Wood, Rub. Chem. and Tech., 49, 189, 1976.
ASTM C177-71, Annual Book of Standards, Part 35, 1, 1976.
D. Hands, RAPRA Members Journal, Dec., 1974.
INDEX

Abrasion
ant ioxidant atmospheres 41 1
brittleness effeas 411
by blade 414
changing atmospheres 41 1
crystalization 408
direction effects 403
energy density at break 411
equilibrium 413
fatigue failure dependence 406, 414
filled rubber 408
humidity effects 414
in nitrogen atmosphere 413
in tires 405
inert atmospheres 411
interpretation of tests 414
load dependence 406
pattern 403, 405
rate 403
roadwear correlation 414
shape dependence 403
start of 405
stick-slip phenomenon 407
strain rate dependence 408
tests for 11
tearing energy dependence 406
temperature dependence 407
tensile strength dependence 406
velocity dependence 407
Accident avoidance testing 556
Acoustical effect, natural frequencies 757
Adhesion
Evaluation 107
compression 112
cord stripping test 109
dynamic tests 110
fabric stripping test 109
flexing 1 14
H.I.T. and U-tests 108
pop-test 108
shearing tests 113
static tests 108
Mechanism 87
Requirements 85
Adhesive friction 309, 711

887
Adhesive treatment
glass fiber 106
nylon 92
polyepoxide pretreated yarn 103
polyester 96
polypropylene 107
rayon 92
steel wire cord 106
vinylon 107
wholly aromatic polyamide 107
Adhesive zone size during cornering 669
Adhesive*:
isocyanates 97
n3 system 104
polyepoxides 98
polyethylene-ureas 97
polyvinylchloride, modified 98
RFL 86, 92
Adsorbed water on dry surfaces 317
Aging test for rubber 8
Air friction 584
Air ratio 711
Aircraft tire
contact pitch 266
effective rolling radius 741
external surface strain. 524
internal surface strain 528
load-deflection curves 732
loading, pressure rise 726
non-steady state motion 835
sidewall force transmission 726
thin shell analysis 490, 493
Amplitude ratio, low frequency input 752
Anchorage, tubular tire 214
Anisotropy, friction 386
Anti-skid materials, ice and snow traction 355
Antisymmetric motion 807
Aspect ratio, rolling resistance 595
Asperities, pressure distribution around 279, 281
Asperity size, rubber friction 386
Assessment of steering characteristics 439
Asymmetric tires 215
Atmospheric effects on abrasion 411
A,V values 400
glass transition temperature effects 402
on ice 402
(rubber compound effects) 402
Auto tire— see passenger car tire
Axisymmetric analysis 516
applications 518
Axisymmetric loads 480

888
Balancing 627
Bead
area properties 880
cable 210
construction 210
cycle tires 210
force measurements 532
forces in 21 1
multiple coil 213
ply wrapping 212
rubber 875
tape 211
wire 209
Belted bias tires 219, 478
cord forces, cornering 531
cord forces, rolling 531
cord load from manufacture 531
cord load in turn up area 531
principal of 303
rolling resistance 594
transmissibility 568
Bending moments 242
Bending stiffness 241, 244
Bending stresses 240
Bias belted tire— see belted bias tire
Bias tire 217, 477
braking and cornering 696
braking force coefficient 606
camber force 688
camber thrust 814
compound properties 893
construction 477
contact patch 270
contact zone cord tension 530
cord angles 485
cord forces and loads, cornering, rolling 531
from inflation, and manufacture 531
critical rotational speed 780, 782
effective radius 582, 744
effective road level 744
enveloping characteristics 732
finite element analysis 503
hydroplaning 327
lateral stiffness 570
load-deflection curves 731
locked wheel braking coefficient 608
longitudinal stiffness 569, 822
models of 791
natural frequencies 761
netting analysis 486
non-steady state simulated response 850
pressure distribution 277
radial run-out 614
resonance 760
rolling resistance 584, 596
rubber additive proportions 875
rubber compounds used 872
sell aligning torque 439
shimmy excitation results 851
side force 439
skid coefficients 444
spring stiffness 566
standing wavelength 780
surface strain 524
transmissibility 568
torsional stiffness 571
turning characteristics 813
vertical axle motion, frequency response 753
vertical stiffness 562
vertical vibration transmission 759
wear during cornering 673
Bonding, rubber to cord 85
adhesive requirements 85
adhesives for rayon, nylon, polyester 86
contributing factors 89
effect of heat treatment 94
methods 86
recent improvement of polyester 105
Braking 308
and cornering, dry road 695
and cornering, wet road 697
carcass construction effects 609
deformation 599
effective radius 606
force-slip relation 608
locked wheel coefficient 608
longitudinal force distribution 601
longitudinal sliding 603
on ice 349
rolling resistance 5%, 606
self aligning torque 697, 815
slip 318
slip ratio 605, 608
traction relations 609
vertical distribution of pressure 603
wet road measurements 606
truck tires vs. passenger tires, wet 712
Braking coefficient 345, 605
contact pressure dependence 709
force curves 700
microtexture dependence 608
peak vs. locked wheel values 608
Breaker cords 217
Breakers
mechanical behavior 219
radial tire function 220
rigid 219
Breaking in of tires 591, 712,
877
Buckling
cord 224
tread 287, 574
Building of tires 225, 229

c
Cable bead •. 210
Calendered sheets 230
Camber
analysis 785
angle
definition 688
lateral force 688
variation 690
antisymmetric motion 807
path curvature 812
pseudo 616
steady state motion 830
slip relations 690
turning combination 810, 814
with sliding 812
Camber force
bias tire 688
cornering 690
correction 690
lateral force ratio 814
pneumatic lead 688
radial tire 688
Camber models 789
Camber moment 688
Camber thrust 814
Carbon black 3, 386
reinforcement of rubber 29
Carcass coat stock
properties 880
rubber 875
Carcass construction
braking 609
loaded radius 582
power loss 594
rolling resistance 594
snow traction 352, 355

891
Carcass flexibility, hydroplaning 344
Carcass stiffness, cornering force stiffness 676
Chafer 214
Chains 351, 353
Characteristics calculated using models 791
Coefficient of braking 345
Coefficient of cornering 345
Coefficient of friction—see friction, coefficient
Compliance, pressure distribution relation 288
Composite slip
assumptions in theory 422
braking torque and turning 425
front wheel drive 427
load transfer effect 427
problems with idealizations 427
self aligning torque 423
side force 423
traction 422
Composite theory
cord rubber
complex modulus 181
failure mechanisms 193
loss properties of composites 189
loss properties of cord and rubber 186
mathematical basis 179
physical models 183
tires
Cough model for tread wear 164
obstacle enveloping 162
ply steer 171
stress analysis 168
tread wear 163
vibrational characteristics 167
Compounds
carcass 4
passenger car tires , 877
rubber 477, 3
snow traction 351
tread 4
truck tire 872
Compression of tread rubber 243
Conditioning of new tires 591, 712,
877
Conicity 616
acceptable levels of 618
direction of 618
effects of 618
inflation pressure effects 618
load effects 618
ply steer effects 617
randomness 617

892
reasons for 618
rim width effects 619
vehicle drift 623
vehicle performance relation 619
vehicle pull 623
Construction of beads 210
Construction methods 225, 477
Contact length
calculations 434
changes due to rough road drift 855
Contact patch
aircraft tires 266
area calculations 262
area-deflection relation 251
area equations 312
bias ply tire 270
cornering 254
developable surface 303
dry area calculation 391
dry road 260
dynamic characteristics 268
effective area 726
elliptical shape 264
environmental effects 255
fluid film 257, 269
force and moment equations 552
force interpretation 552
force transformations 552
friction 316
gross area 269
hydroplaning 326
ice traction 349
isotropic solid elastic bodies 262
length-deflection relation 728
loss of 257, 269
non-developable surfaces 303
on and off road tires 261
passenger car tire comparisons 269
pressure distribution 273
shape 251, 268
slip 293, 303,
416
sphere-road equation 222
strain 222
thin film 346
theory 261
toroid-road equation 221 , 303
tread compression 243
tread deflection 574
velocity dependence 253, 270
with curved surfaces 253

893
zone concept 257
Contact patch stresses 243
braking relations 296, 300,
302
cornering force relations 297
friction coefficient relation 293, 300,
302
idealizations 292
lateral component 296
longitudinal component 294, 302
measurements 295
separation of causes 2%
shear 294
stationary tires 294
summary of 302
tangential component 292, 303
traction relations 296, 300,
302
yaw relations 297
Cord
adhesion mechanism 87, 108
breaker 217
buckling 224
calendered sheets 230
comparative analysis 65
crowding 231, 236
design 208
effects of twist 80
elastic constants 129
fatigue resistance 73
fiber glass 61
final path 235
flatspotting mechanism 51
impact resistance 68
length calculation 321
materials 207
movement during manufacture 234
nylon 43
other less known types 63
path considerations 229
physical properties 38, 62, 65
polyamides, wholly aromatic 63
polyester 55
pre-rubbered sheets 230
rayon 39
steel 60
steel wire for truck tires 599
Cord angle
bias tire 485
rolling resistance 594

894
Cord forces
cornering 531
radial tire 516
rolling tire 531
Cord loads 531
measurement 529
Cord path analysis 235
Cord reinforced rubber, failure mechanisms 193
fatigue 194
filament failure 195
Cord rubber composites, loss characteristics 179
measurement methods 184
Cord stiffening 489
Cord stripping test 109
Cord tension 243, 482,
488
bias tire 530
Cordless tires 65, 479
Coriolis force 748
Cornering
adhesive zone, size dependence 669
bias tire 696
braking 695
contact patch 254, 297
curve shifting 676
deformation velocity 669
emergency vehicle handling 674
irrecoverability of lateral force 678
lateral force distribution 670
literature references 785
moment 671, 676
parameters affecting 674
radial tires 696
rolling resistance 5%
sliding velocity 669
slip angle effects 676
steady state motion 829
tread rubber stiffness dependence 676
vertical load variations 852
wet roads 677
Cornering coefficient 345
Cornering experiments 664
combination deformation 669
Cough apparatus, use of 665
lateral deformation 665, 674
longitudinal deformation 668
low slip angles 668
movable platform method 665
photographic method 664

895
Cornering force
braking effects 81J
camber effects 688, 690
coefficient 700
coefficient changes 699
curves 673
measurement 692
model 793
oversteer ". ."7 697
tests
conventional vehicle 701
drum vs. road 715
equipment limitations 713
non-tethered traction test 701
reliability of data 701
slalom test 701
tire conditioning 712
traction effects 815
traction-slip angle relation 697
understeer 697
vanishing sliding 805
variations due to rough road drift 860
wet roads
coefficient 685
drainage effects 680
footprint dependence 680
hydroplaning 687
inflation pressure dependence 684
load dependence 680, 682
load transfer effects 682
radial tires 698
surface dependence 687
traction 700
tread effects 680, 685
truck tires 683
vehicle handling 680
velocity dependence 685
wear dependence 685
Cornering moment stiffness 676
Cornering power of tire 423
Cornering stiffness
carcass elasticity model 822
elastic tread element model 793
elasticity model without carcass 819
simple stretched string model 824
vanishing sliding 807
Cost of tires 872
Cracking of rubber 11
Creep 785
Creep velocity in tire models 816
Critical rotational speed 779

896
factors affecting 782
Cross ply tires—see bias tire
Crown
angle 217
overlap 212
Crystallization of rubber surface 385, 408
Cure, rubber 7, 485
Curing bag 229
Cycle tire
bead 210
manufacture 229
Cylindrical former 237

D
Damping
frictional 312
longitudinal 754
Defects, pressure distribution 276
Deflection
contact area relation 251
footprint loads 509, 519
free rolling 743
Deformation
bending 287, 293
braking 599
interlaminar 157
lateral 658
membrane 293, 303
radial 286
running band 286
shear 287
shell structure 223
slip 744
tread 300
Deformation velocity 669
Delft drainage meter 706
Demounting of tires 214
Design
for thick fluid films 324
of cords 208
of filaments 224
of tread elements 260
problems 207
tradeoffs 880
Displacement
fields 513
functions 499, 514
matrix 500
Drainage
air ratio 711

897
measurement 704
meters 677, 706
passenger car tires 709
shore length ratio 711
tire characteristics 709
truck tires 708
Drift 785
ply steer 623
rough road 853
vehicle 620, 623
Driver control
during hydroplaning 328
on ice 349
safety 438
steering information 420
vehicle combination 542
Drum tests
external 546
internal 540
rolling resistance 591
vs. road tests 715
Dry contact 260
Dry friction 309, 316
Dry pavement 322
Dry surface, adsorbed water and viscosity 317
Dual tires, wear 599
Dynacor rayon 41
Dynamic loading 480
Dynamic modulus of rubber 22, 25
Dynamic pressure changes 577
Dynamic tests of rubber 11

E
Effective deflection 581
Effective radius 581
Effective road level 581
Effective rolling radius 577
Elastic constants
cord 129
multi-ply systems 146
one ply systems 130, 141
rubber 126
Elastic instability of rubber 381
Elasticity of rubber, theory
molecule 15
molecular network 17
strain energy representation 19
thermodynamic 13
Electrical conductivity of rubber 386
Emergency maneuvering 674

898
Energy loss
composites 189
cord and rubber 186
rolling resistance 588, 596
tires 27
Energy storage in rubber 876
Enveloping of obstacles 727, 732
circumscribed circle shrinkage 733, 736
composite theory 162
drag force of 737
experiment-theory agreement 736
force variation equations 739
force variation—obstacle height relation 738
foundation stiffness nonlinearity 736
height to contact length relation 736
inflation pressure effects 733
longitudinal force variation relation 737
pressure distribution 737
road contour effects 738
softening foundation 733
theory 733
unit bump response 740
unit road step response 738
vehicle response to obstacle 741
vertical force variations 732, 735,
737
Environmental effects in contact patch 255
Equations of rolling and slipping 786
Equilibrium equations for thin shell analysis 490
Equilibrium shape 241, 485
Experimental techniques 522

F
Fabric stripping test 109
flaw correlation 637
Failure
mechanisms 193
starting points 247
Fatigue
failure of tires 194
life 212, 247
resistance of cord 73
tests 78
Fiberglass cord 61
Filament 224
Filled rubber compounds 386
Finite element analysis 247, 4%
computer use for solution 498
constant strain triangular element method 499
displacement methods 498

899
force methods 498
gap element method 522
hexahedral element method 512
history 497
hybrid methods 498
improvements in methods 511
isoparametric elements 511
methods, choice of 498
non-axisymmetric analysis 504
shell element methods 508
subparametric element analysis 512
Flange height, effects on rim-tire interference 215
Flatspotting mechanism 51
Flaw
failure correlation 637
location of 636
Flex cracking of rubber 11
Flexible membranes 237
Flexing test , 1 14
Fluid film
contact patch 257, 269
friction 315
hydrodynamic lubrication 391
pressure distributions 277
regions 322
thick. . . . , 322
thin 260
traction 270, 322
viscosity effects 255
Flutter 629
Footprint loads 509, 519
Force
bead 211
measurement, bead area 532
radial 246
resolution 243
transmission, side-wall 726
variations, neglecting inertia 749
Former, cylindrical 237
Free rolling 743
Frequency
rotational natural frequency 752
rough road tire drift 860
rubber properties 876
Frequency response
longitudinal axle motion 754
non-steady state motion 839
radial run out 752
road irregularities 757
static deflection variation 756
vertical axle motion 751

900
Friction
adhesional energies 381, 384
adhesive 309, 316,
391, 711
air 384
angular relations 308
anisotropy 386
asperity size dependence 368
bond theory and failure 381
boundary lubrication 388
buckling 370, 383
coefficient 308, 367,
377
compound dependence 396
construction dependence 702
contact pressure dependence 470
effective lateral values 815
effective values 434
film thickness effects 391
ice 394
inflation pressure dependence 702
load dependence 367, 394,
702
low slip 419
measurement difficulties 704
macrotexture dependence 704
microtexture dependence 711
peak values 393
polymer blends 389
skidding 393
sliding 607
surface dependence 434
temperature dependence 310, 367,
389, 401,
702
velocity dependence 310, 367,
381,389,
392,401,
702
components in tire 309
compound dependence 3%
control forces 308
curves 377, 388
filler transformation 397
ice transformation 394
damping 312
deformation 312
dry 309, 316,
391,711
electrical conductivity relation 386
equations for prediction and validity 371

901
equilibrium values 385
experimental methods 369
filler effects 3%
fluid films 315
free rolling 308
frequency dependence 312
glass transition temperature dependence 379, 394,
398
hard sliders 370, 383
heating 311
hydrodynamic lubrication 388, 391
ice 394
induced vibration 310
large slip velocities 317
lift 373
load dependence 367, 394,
702
loss modulus relation 379
maximum values 379, 393
measurement 370
membrane thickness 318
microtexture dependence 315
oil extension 371, 397
peak 379, 393
prediction 310
polar substance effects 388
property dependence 317
road surface effects 313
rolling 380
rough track 368
skid coefficient 393
stick-slip process 380
surface structure effects 311
tangential forces 308
tangential strain 380
tearing 316
temperature effects 314, 367,
375, 389,
394, 401,
702
testing 379
theory-experiment agreement 369
tractive forces 308
tread element orientation 679
velocity dependence 310, 316,
367, 392,
381, 389,
401, 702
visco-elastic process 377
viscous 315
wave mechanism of sliding 383

902
waves of detachment 382
wet 314, 390
Front wheel drive 427
Functions
tire 204
tread 216
Future analysis, developments 522

G
Gap elements 522
Geodetic movements 232
Geometry
changes due to inflation 240
changes due to loading 480
definitions 482
description of tires 220
predictions 487
Glass fiber, adhesive treatment 106
Glass transition temperature 379
Gough model for tread wear 164
Grinding
correction 634
disadvantages of 635
eccentric machining 633
force correction 633
non-uniformity correction 633
on-car grinders 634
roughness 634
shake 634
truing 633
Grooving of roadway 327
Gyroscopic couple 845
Gyroscopic effect, non-steady state motion 835

H
H-test, cord adhesion 108
Halpin-Tsai equations. 131
Hardness, rubber 10
Harshness 630
Hexahedral element analysis 511
High speed tires 217
History
analysis 872
building tires 229
finite element methods 497
force measurement 546
models 481
moment measurement 546
stress analysis 523
tire structures 477

903
Holographic interferometry 527
Holography, tire inspection 637
Hooke's law
isotropic materials 128
orthotropic materials 137
Hoop tension forces 729
Hop, wheel 630
Horizontal load-deflection relation 568
Hybrid simulation 825
Hydrodynamic lubrication 388, 391,
401
film thickness effects 391
directional effects » 434
Hydroplaning
bias tires 327
carcass flexibility effects 344
contact length dependence 326
cornering force 687
driver dependence 328
explanations 339
grooving of roadway 327
load dependence 325
porous roadways 328
pressure distributions 277
radial tires 327
steering feel 687
testers 546
tread dependence 325
tread reversion 338
velocity dependence 324
viscous 338, 346
Hysteresis, rolling tire 727
Hysteresis loop 562

I
1-test 108
Ice friction
braking temperature rise 349
coefficient 394
contact area effects 349
filler effects 396
glass transition temperature dependence 394, 398
oil extension effects 397
tire design for 349
Impact resistance of tire cord 68
Impractical designs for tires 205
Inertia effects 726
Inflation
load carrying 726
loading 245, 479
pressure distribution 273
shape changes 240
Inflation pressure
coefficient of friction 702
conicity . 618
cornering force , 680
dynamic changes 577
enveloping of obstacles 733
lateral stiffness 684
ply steer 619
radial stiffness 562
self aligning torque 684
thin film traction relation 341
traction 329
Inspection, tire
holography 637
optical 637
ultra-sound 645
visual-tactile 636
x-ray 639
Interfacial bonding 311
Interference, rim-tire 214
Interferometry 527
Interlaminar deformations 157
Internal strain measurement 527
Interply load distribution 522
Interply shear strain measurement 528
Isocyanates 97
Isoparametric element analysis 511
applications 516
deformation analysis 516
hexahedral elements 512
inflation stress analysis 516
limitations of use 522
subparametric elements 512
three dimensional analysis 512
two dimensional analysis 515
use with variable properties 515

K
Kerfs 274, 337,
346
truck tire wear 685
Kinetic energy of tire ^ 766
Kinematic shimmy 842

L
Laminate stiffness 491
Lateral behavior of tire models 791
deflection at vanishing sliding 805
Lateral deflection 658, 668,
674

905
changes from rough road drift 854
Lateral force
amplitude ratio 674
camber angle relation 688
camber force ratio 814
distribution during cornering 670
measurement techniques 671
non-steered vehicles 658
steered vehicles 661
transmission 762
Lateral foundation stiffness 729
Lateral runout 632
Lateral sliding of tread 666
Lateral stiffness 569
bias tire 570
foundation characteristic effects 729
inflation pressure effects 684
radial tire 570
test method 570
vertical stiffness relation 571
Lateral stresses 296
limits of calculation 290
Lateral vibration
BOhm theory 770
experimental method 770
theory-experiment agreement 772
Lift, frictional 373
Load
abrasion 406
bias tire deflection curves 731
carrying
actual vs. theory 244
addition of parts (superposition) 244, 479
analysis 247
inextensible membrane theory 238
inflation pressure 245
non-ideal considerations 238
pneumatic 206, 237
simplifications for theory 243
spherical membrane theory 238
structural 247
transition curve 239
transmission 245
uninflated 241
coefficient of friction 394
conicity 618
cornering force 682
deflection theory 727
experimental agreement 731
footprint area 509, 519
horizontal relations 568

906
lateral stiffness effects 730
linearity 566
membrane concept 728
radial deflection relation 730
radial stiffness effects 730
reduced load-deflection curves 731
tread band rigidity effects 729
vertical relations 560
effective radius 583
hydroplaning relation 325
interply distribution ' 522
ply steer 618
power loss 592
rolling resistance 592
rubber friction 367, 394
self aligning torque 422, 683
slip equivelance 421
steady state motion 829
transfer during turning 246, 427,
682
variation due to road roughness 852
wear 457, 462
wet traction 331
Loaded radius 577
braking effects 584
carcass construction effects 582
inflation pressure effects 583
tractive effects 584
tread thickness effects 583
velocity effects 584
Loading
axisymmetric 480
bending stresses 240
dynamic 480
geometric changes 480
inflation 479
inflation pressure rise 726
non-axisymmetric 480
response 480
shear 240
static 480
thermal 480
Locked wheel braking coefficient 608
Longitudinal axle motion frequency response 754
Longitudinal damping ratio 754
Longitudinal force
angle dependence 818
brush type tire model 828
Coriolis force 749
distribution during braking 601
nonuniformity effects 746

907
over-road irregularities 746
steady state motion 830
theory, development 747
transmission 762
variable parameter effects 747
Longitudinal response, bias tire 760
Longitudinal sliding, braking 603
Longitudinal slip 741
causes 744
contact length dependence 745
deflection dependence 745
large slip values 746
ratio 816
stiffness 745
values 742
velocity effects 742
Longitudinal stiffness 568
bias tire 569
model with carcass elasticity 822
radial tire 569
test methods 568
Longitudinal stresses 294, 302
Loss properties
composites 189
cord and rubber 186
Low frequency response, amplitude ratio 752
Lubrication
boundary 388
directional hydrodynamic effects 434
polar substances, effects of 388
hydrodynamic 388, 391

M
Macrotexture 258, 312,
704
braking force coefficient 704
coefficient of friction 704
drainage 335, 704
measurement 335, 344
profile ratio 704
roughness meter 704
skid resistance 333
texture depth 705
thick film traction 333
void width 704
Manufacturing methods 226
problems 234
Mapping 267
Matching, tire-rim 633

908
Measurement
bead forces 532
braking deformation 600
contact patch stresses 295, 554
cord forces 529
cornering force 692
cornering deformation 665, 668
difficulties 533
drainage 677
drum methods 591
effective rolling radius 581
forces and moments 546
future developments 534
horizontal load-deflection relations 568
internal strain 527
interply shear 528
lateral forces 671, 673
lateral runout 613
lateral stiffness 570
longitudinal forces 601
longitudinal sliding 603
macrotexture 740
microtexture 711
moving belt machines 590
nonuniformity 61.3, 632
radial runout 613
radial stiffness 562
rolling resistance 588
roughness 704
skid resistance 677
slip angles 692
surface strain 523
techniques 542
temperature 597
torsional stiffness 571
trailer usage 594
tread deflection 574
turn-up area 531
vertical dynamic forces 573
Mechanical interferometry 526
Membrane
analysis 481, 489
basis for theory 268
concept for load-deflection theory 728
deflections 241
deformation 293
flexible 237, 338
inextensible 238, 253
tension forces 241
thin 237, 338
toroidal 240

909
Microtexture 258, 312,
315
classification of 711
coefficient of friction 315, 71 1
ice 349
importance 340
measurement 711
penetration of 340
rubber rupture 711
skid resistance 448
thick film traction 335
Models
anisotropic ply 480
comparisons of 791
cord and rubber 480
equivalent structures 481
finite element 481, 4%
for rough road tire drift 853
gap elements 520
history 481
isoparametric element 511
laminate models 481
lateral behavior 791
membrane shell models 481, 504
netting analysis 481
non-axisymmetric analysis 481 , 504
plane-strain models 519
shell elements 508
small scale testing rules 851
steady state motion 790
stretched string 796
thin shell analysis 481, 490
with carcass elasticity 820
with elastic tread 793
without carcass elasticity 816
Modulus of rubber 5
Moire fringe method 526
Molding 225, 229
Moments, bending 242
Miihlfeld antenna system 574
Multiple bead coils 213

N
Natural frequencies 757
bending stiffness effects 770
Bbhm theory and calculations 769
experimental methods 770
node positions 769
Netting analysis 481
bias tires 486
calculation problems 489

910
equilibrium shape 485
force resolution 484
^limitations of 488
membrane forces 482, 287
other uses of 485
pantographing 485
radial tires 489
radius of curvature 487
Neutral steer 660
Non-axysymmetric analysis 504
Non-steady state motion 833
aircraft tires 835
approximations used in theory 842
beam concept 834
computer use in calculation 843
contact equations 835
distance dependence 863
finite slip values 862
forces acting on tire 850
frequency response functions 840
gyroscopic effect 835
history of investigation 834
horizontal motion 833
influence of tire mass 847
interaction of vertical and horizontal 863
load-lateral force relation 863
low slip values 853
model usage 833
Nyquist curves 850
partial sliding behavior 845
response to sinusoidal input 839
response to step functions 839
rough road tire drift *• 853
string concept use 834
theory development 849
theory-experiment agreement 868
time varying load effects 852
transfer functions 835
tread pattern effects 842
Non-steady state tire models 790
Non-tethered traction test 701
Non-uniformity 612
conicity 616
contact measurement methods 613
contactless measurement methods 614
corrections for 632
dimensional 613
effects of 629
grading machines 63 1
lateral force variation 616
longitudinal force variation 624, 746

911
other measurement techniques 632
ply steer 617
radial force variations t615
reasons for 613
testing conditions 632
unbalance 627
vibrations 630
wheel 627
Normal pressure shifts ' 254
Nylon tire cord
adhesives for 86
heat aging resistance 44
heat treatment 45
high tenacity 43
production 44
strength loss 47
treatment for adhesives 92
Nyquist curves 850

o
Off-road tire 356
contact patch 261
Operational severity 335
Optical inspection of tires 637
Overlap, crown 212
Oversteer 660
cornering force 697
Oxidation of rubber 875
Ozone attack of rubber 875

P

Pantographing
movements 232
reliability 236
use 485
Passenger car tire— see also belted bias, bias, or radial tire
contact patch 251, 268
pressure distribution 292
rubber compounds used 872
sidewall force transmission 726
slip 305
Pavement structures 258
see macrotexture or microtexture
Pendulum skid tester 451
Performance rating of tires 341
Permeability of rubber 875
Phase angles
self aligning torque 435, 440
side force 435, 440
Pitch and yaw correlation 718

912
Plane-strain models 519
Platform tire testers 548, 558
Ply
arrangement 217
description 477
slippage 235
steer
acceptable levels 618
avoidance in design '. 618
composite theory 171
conicity effects 617
extreme testing methods 623
inflation pressure effects 619
load effects 618
magnitude of 617
reason for 618
rim-width effects 619
vehicle drift 623
vehicle performance 619
wrapping 212
Pneumatic lead 688
Pneumatic load carrying 237
Pneumatic pressure 243
Pneumatic trail 298
model with elastic tread 794
model without carcass elasticity 820
steer angle relation 298, 658
vanishing sliding 807
Poissons ratio, rubber 877
Polyamide tire cord 63
Polyepoxide
adhesive 98
pretreated yarn 103
Polyester
adhesive for 86
adhesive treatment 96
cord 55, 57
tire durability 58
Polyethyleneureas adhesive 97
Polypropylene adhesive treatment 107
Polyvinylchloride adhesive ,. 98
Pop test, cord adhesion 108
Portable skid resistance meter 677
Potential energy of tires 765
Power loss of tire 592
Pressure distribution
bending theory relation 292
bias tire 277
compliance relation 288
contact patch 273
defining variables 274, 286

913
flooded tread 277
inflation pressure relation 273
interfacial 292
mathematical models 289
normal component 271
passenger car tires 292
pavement structure dependence 279
roadwheel 283
shear 272
soil and sand 282
structural parameters -. 282
tangential component 271
tire defects 276
tread rubber hardness effects 281
velocity dependence 277, 291
vertical components 299
Pressure shifts 254
Pressure spikes 274, 287
Profile
families 488
ratio 704
Properties, rubber—see rubber
Pseudo camber 616
Pseudo slip 617
Pull 619

R
Radial deflection ' 730
Radial deformation 280
Radial foundation characteristics 728
Radial forces 246
Radial passenger car tire compound properties 881
Radial runout
acceptable standards 632
bias tire 614
conclusions from 614
curves 614
frequency response 756
radial tire 614
Radial stiffness
inflation pressure dependence 562
measurement methods 566, 573
theory 727
Radial tire
braking and cornering combination ' 696 i
braking force coefficient ffif^iOf
belt edge strain 526, 529
breakers 220
camber force 688
camber thrust 814
compounds 872

914
cord forces 516
cord loads from manufacture 531
critical rotational speed 782
deformation due to inflation 516
description 219, 478
effective radius 582
effective road level 744
effective rolling radius 744
enveloping characteristics 732
force transmissibility ,.. 568
history 220
hydroplaning 327
instability 220
lateral stiffness 570
load-deflection curve 731
locked wheel braking coefficient 608
longitudinal response 760
longitudinal stiffness 569, 822
model for analysis 791
natural frequencies 761
netting analysis 489
non-steady state simulated response 850
principal of 303
radial runout 614
resonances 760
rolling resistance 588, 594
596
rubber additive proportions 875
self-aligning torque 433, 439
shell structure 286
side force 433, 439
sidewall strain 528
skid coefficients 444
snow traction 352
surface strain 524
torsional stiffness 571
turning characteristics 813
vertical axle motion, frequency response 753
vertical stiffness 562
vertical vibration transmission 759
vibration characteristics and composite theory 167
wear 460
during cornering 673
wet traction 337
Radial truck tire compound properties 882
Radius
effective rolling 577
loaded 577
of curvatures 239, 486
transformation 232

915
Rayon tire cord
adhesive;, for 86
Dynacor 41
extra high modulus 42
high tenacity 39
Reinforcement of rubber, carbon black 29
Relaxation length 569
simple stretched string model 824
stretched string model with tread 802
vanishing sliding 806
Relaxation time 675
Resilience of rubber 455
Resonance 760
Reversion of rubber 6
RFL adhesive 86
improvements 95
pickup 93
treatment of nylon and rayon 92
Rigid breakers 219
Rim-tire combination
flange height effects 215
interference 214
matching 633
reactions 241
taper effects 214
types 214
Road
contour decomposition 738
curvature-stiffness relation 727
effective level 744
irregularity frequency response 750
roar 759
tire contact 250
Roadway
antiskid materials for ice and snow 355
grooving of 327
hydroplaning relation. 327
macrotexture 258
microtexture 258
porous 328
surfaces 250
Roadwheel pressure distribution • 283
Rolling body equations 786
Rolling, definition 578, 785
Rolling radius
aircraft tire 741
effective * 577
velocity dependent 741

916
Rolling resistance 28, 584,
727, 741
aspect ratio effects 595
belted bias tire 594
bias tire 594, 596,
744
braking 596, 606
carcass construction effects 594
coefficient 588
cord angle effects 595
cornering effects 5%
energy loss of 588, 596
indoor tests 590
inflation pressure dependence 592
load dependence 592
longitudinal component 741
measurement methods 588
moment 750
number of ply effects 595
radial tires 588, 594,
596
rim width effects 595
road surface effects 589
road tests 588
rubber and cord compound effects 594
slippage losses 596
temperature dependence 593
tire conditioning effects 591
traction effects 5%, 606
truck tires 595, 598
units of 587
velocity dependence 588, 592
worn tires 594
Rolling velocity vector 787
Rotational natural frequency 752
Rough road tire drift 853
amplitude effects 860
comparison of models 858
frequency effects 860
models for 853, 857
self aligning torque 858
side force 858
Roughness 630
Rubber
additive proportions 875
aging tests 8
air permeability 875
as engineering material 875
bead area 875
bias tire compounds 872
bonding methods 86

917
carbon black filling 386
carcass coat stock 875
compounds 477
for snow traction 351
cord bonding 85
cure 7
dynamic modulus 22, 25
dynamic properties 878
dynamic tests 11
elastic constant 126
elastic deformability 875
electrical conductivity 386
energy storage 876
flexibility for snow traction 351
frequency dependence of properties 876
friction 317, 367
glass transition temperature 379, 394,
398
hardness 10
oxidation 875
Poissons ratio 877
recoverability 876
rolling resistance 876
rupture 92, 340
shear 309
shear modulus 878
sidewall 875
strength and modulus 875
strain energy functions 19
tear tests 9
temperature dependence of properties 876
thermal conductivity 879
time dependence of properties 876
viscoelasticity 20
wettability 345, 351
yield point 875
Youngs modulus 5, 126,
386, 877
Running band deformation 286
Runout
lateral 613, 632
radial 613
Rupture, tire
fiber layer '. 91
rubber layer 92

S
Salted roads 355
Sand-soil pressure distribution 282
S. C. meter, slip and camber meter 694, 718
Scrub, wear 690

918
Self aligning torque
bias tire 439
braking effects 697, 815
brush type tire model 828
composite slip 428
constructional details 433, 439
contributing factors 791
deformation due to 849
equations for calculation 429
equilibrium 443
frequency dependence 432
high slip angles 677
inflation pressure 684
load dependence 435, 683
low slip angles 676
maximum values 435
measurement, drum vs. road tests 715
model with carcass elasticity 821
model with elastic tread elements 793
model without carcass elasticity 817
"non-linearity 432
non-steady state motion 837
phase angles 435, 440
radial tire 433, 439
reduction of slip angle 420
road condition influences 431
rough road tire drift 858
slip angle 658
steady state motion 829
time delays 428
theory-experiment comparison 430
traction effects 697, 815
transformation to universal curves 433
tread effects 440
vanishing sliding 805
velocity dependence 432
wet roads 420
Shake 629
Shape
changes due to inflation 240
equilibrium 243
Shaping, tread 225
Shear
loading 240
modulus of rubber 878
pressure distribution 271
rubber 309
Shell
bending models 253
structure 286
structure deformation : 223

919
Shell element analysis 508
Shimmy 629, 842
bias tire excitation 851
Shore length ratio 711
Side force
bias tire 439
brush type tire model 428
causes 658
composite slip 423
constructional detail 433, 440
deformation due to 849
equations for calculation 429
equilibrium 433
frequency dependence 432
load dependence 435
maximum values 435
model comparisons 791
models used 790
model without carcass elasticity 817
non-linearity 432
non-steady state motion 837
phase angles 435, 440,
444
radial tire 433
road condition influences 431
rough road tire drift 858
sloped roadways 661
steady state motion 832
theory-experiment agreement 430
time delays 428
transformation to universal curve 433
transient motion 661
tread effects 440
velocity dependence 432
wind effects 660
Sidewall
force transmission 726
properties 880
rubber 875
strain 528
tension 246
white sidewalls 875
Six component tire tester 550
Skid
coefficient 442
velocity dependence 393
description 416
distance equation 341
numbers 712
resistance
compound effects 451

920
construction effects 449
front wheel braking test 441
locked wheel trailer test 442
macrotexture dependence 333
meters 677
microtexture dependence 448
ranking on ice and water 451
road surface effects 447
rolling resistance effects 442
siping 449
tests 441
theory-experiment agreement 446
tread pattern effects 446
values 712
velocity dependence 444
wet roads 712
temperature rise 400
testers 711
reliability 451
velocity dependence 400
wet roads 712
slalom test 701
Sliding
friction coefficient 607
temperature increase 465
wetness effects 466
velocity vectors 787
zones
turning 812
vertical force dependence 671
Slip
acceleration effects 416
angle 785
camber effects 690
centrifugal force relation 661
correction charts 694
measurement 692
model without carcass elasticity 817
self aligning torque 658
steady state motion 832
temperature increase 463
toe in 690
assumptions, theoretical 422
belt analogy 307
braking effects 318, 416
calculation 418
circumferential 419
composite 422
contact patch 293, 303,
416
strain 417

921
deformation 744
energy loss 416
force components 419
friction effects 309
longitudinal free rolling 742
measurement 305
passenger car tires 305
primary 304
property effects 309
pseudo 617
ratio
braking 605, 608
traction 610
secondary 303
self aligning torque, time delay 428
side 319. 417
side force, time delay 428
simple 416
stiffness 744
stiffness effects 418
straight line motion 304
tangential force relation 305
tangential traction 418
tread compound dependent 303
turning 318
velocity
brush type tire model 823
tangential 741
yaw 304
Slip and camber meter 694, 718
Slippage between plies 235
Slipping body equations 786
Snow traction 350
Softening foundation characteristic 733
Sphere-road contact patch 222
Spherical load carrying 238
Spikes, pressure 274, 287
Spin 785, 807,
810
Spokes 241
Spring rate of tires 566
Standing waves 727, 773
bias tire 780
critical rotational speed 779
damping 773
deformation 773
drum tests 782
flexural rigidity influences 778
formation of 774
heat buildup 773
membrane theory applied to 773

922
power consumption of 773
propagation 774
theories 773
experimental agreement 782
Static deflection
force variations 750
variation frequency response 756
Static loads 480
Steady state motion
antisymmetric 807
camber effects 830
cornering 829
load effects 829
longitudinal force effects 830
self aligning torque and side force 829
side slip and camber 829
slip angle calculation 832
tire models for 740
vehicle dynamic behavior 833
Steel wire cord 60
adhesive treatment 106
Steer
angle, pneumatic trail relation 298
neutral 660
Steered vehicle 661
Steering
characteristic assessment 439
driver information 420
Stick slip 311, 380.
702
abrasion 407
brush type tire model 825
Stiffness
bending 241, 244
calculation 434
laminate effects 491
lateral 569
limits of 461
longitudinal 568
slip effects 745
radial 562, 573
radial tire 822
road curvature relation 727
torsional 571
vertical 562
lateral relation 571
Strain
abrasion rate 408
acceleration effects 418
braking effects 417
contact patch 222, 417

923
energy functions, rubber 19
field interference 345
measurement 523, 527
rate, abrasion 408
rubber, energy functions 19
Stress
analysis, composite theory 168
analysis history 523
braking effects 2%
contact patch 243, 292,
302
determination 247
-strain, rubber 5
traction effects 2%
Stretched string tire models
with tread elements 7%
without tread elements 803
Structural
analysis 477
load carrying 247
pressure distribution relation 282
requirements of tire 204
rigidity 562
stiffness 244
Studs 351, 353
Subparametric element analysis 512
Superposition principal 23
Surface strain measurement 523
Suspension, vehicle vibration 560

T
T-test 108
Tangential
force-slip relation 305
force variation 750
slip values 749
slip speed 750
strain 380
stresses 292, 303
Tape beads 21 1
Taper, rim 214
Tearing
friction 316
tests 9
Temperature
abrasion 407
coefficient of friction 310, 367,
389, 394,
401, 702
durability relations 598
energy relations 597

924
measurement 597
power loss 593
rise 593
during skidding 400
rolling resistance 593
rubber properties 876
wear 462
Tenacity, tire cord
nylon 43
rayon 39
Tension forces 241
cord 243
sidewall 246
Thermal
conductivity of rubber 879
loads 480
Thick fluid film 322
traction 344
Thin fluid film
conditions for 345
tire-road interface 346
traction
construction variables 336
macrotexture dependent 344
slip resistance 345
tread pattern effects 337
viscosity 315
Thin membrane analysis 237
Thin shell analysis 490
Thump 629
Time delays
self aligning torque and side force 428
steering 439
Tire
aircraft 266
analysis history 872
analyzer 637
asymmetric 215
belted bias 219, 478
bias ply 270, 477
break-in 877
building 225
characteristics 250
dynamic behavior 726
enveloping of obstacles 727, 732
low frequency properties 726
test methods 715
complexity 542
construction 477
contact patch 250
aircraft tire 266

925
bias tire 270
off road tire 261
cord 207
cordless 65, 479
cost 872
cycle 229
demounting 214
design tradeoffs 880
flaw-failure correlation 637
flaw location 636
functions 204, 542
geometric descriptions 220
high speed 217
inspection 636, 645
instability 220
mapping 267
matching
rim 626
wheel 633
materials 223
models
brush type 825
comparison 792
elastic tread with rigid carcass 793
hybrid 825
non-steady state 790
small scale testing rules 851
steady state 790
stretched string with tread 7%
with carcass elasticity 820
without carcass elasticity 816
molding 225, 229
motion definitions 785
non-uniformity 613, 624,
627
profile 482
property modification 477
radial ply 219, 478
-rim interference 214
structural requirements 204
testing 546, 549.
637
tubeless 209
tubular 209
vulcanization 225, 477
wear, see wear
Toe in 690
Toroid road contact patch 220
Toroidal membranes .240
Torsional stiffness 571

926
Traction
braking relation 609
composite slip 422
construction variables 612
cornering 700
fluid film effects 322
force coefficient 610
frict ional processes 308
hysteretic 335
ice 349
inflation pressure dependence 329
load dependence 331
off-road tires 356
partial contact 328
pavement structure dependence 333
rating 342
rolling resistance 596, 606
shear force distribution 609
slip ratio 610
snow 350
strain field interference 345
temperature rise 401
thin film 336
. tread
compound dependence 330, 612
compression 609
depth dependence 331
design dependence 329, 612
velocity dependence 331
wet
complexity of problem 332
cornering 700
dry comparison 255, 270
factors affecting 322
measurements 610
tread pattern effects 341
Tramp 630
Trailers as measurement devices 442, 589
Transmissibility 568
Transient properties 556, 715
Tread
additives 351, 354
buckling 287, 574
compounds 4, 353,
880
compression 609
cornering force, rubber stiffness 676
deflection, contact patch 574
deformation
braking 600
tire models 816

927
design 260
snow traction 350, 353
wet traction 329
flexible element 270
functions 216
hydroplaning 325
lateral deflection 658
lateral sliding ' 666
materials 216
pattern 216, 274
cornering force 685
thin film traction 337
pressure distribution 281
shaping 225
slippage 584
wear 164, 673
Tread band
free vibration 764, 768
potential energy 765
strain 765
Truck tires 595, 598
Truing 633
Tubeless tires 209
truck tires £99
Tubular tires 209
anchorage 214
Turning
asymmetric motion 807
camber combinations 810
composite slip 425
forces 662
sliding zones 812
slip 318
spin effect 813

u
U-test 108
Ultimate performance rating 341
Ultrasound 645
Unbalance 627
Understeer 660, 697
Undertread properties 880
Uninflated load carrying 241

V
Vehicle
drift 620
dynamics, steady state motion 833
handling, emergency maneuvers 764
pull 619

928
response tests 716
response to obstacles 741
tuning 762
Velocity
abrasion 407
coefficient of friction, influence of 310, 392,
401, 702
contact patch 233, 270,
367. 381,
389
cornering force 685
effective radius 584
hydroplaning 324
loaded radius 584
longitudinal damping ratio 755
longitudinal slip 742
power loss 592
pressure distribution 277, 291
rolling radius 741
rolling resistance 588, 592
self aligning torque 432
side force 432
skid resistance 444
skidding 400
traction 331
vertical stiffness 562
Vertical axle motion, frequency response 751
Vertical dynamic stiffness 573
Vertical force
distribution during braking 603
sliding zone dependence 671
transmission 761
Vertical load-deflection relation 560
Vertical spring rate 566
Vertical stiffness %
amplitude effects 727
bias tire 562
lateral stiffness relation 571
radial tire 562
rolling effects 727
velocity dependence 562
Vertical vibration
flexibility dependence 562
node movement 761
sidewall displacement 761
transmission 560, 757
Vibration
filtering of 556, 558.
631
friction induced 311
intensity, natural frequencies 757

929
nonuniformities 612, 630
radial tire characteristics 167
standing (non-rolling) 311
suspension dependence 560
tire testers for 556
transmission 757
Vinylon adhesive treatment 107
Viscoelasticity of rubber 20
Viscous friction 315
hydroplaning 338, 346
Vulcanization 225, 477

W
Waddle 630. 694
Wall tensions 241
Waves of detachment 382
Wear
bias tire 673
composite theory 163
construction effects 459
cornering 673, 685
cornering power dependence 459
dual tires 599
energy consumption 471
equations for 465
experiment-theory agreement 458
frictional energy dissipation 456
laboratory studies 673
lateral force transmission effects 468
lateral stiffness effects 460, 676
load dependence 457, 462
maximum values 457
pressure dependence 455
pressure distribution effects 467, 470
radial tires 460, 673
rate of 455
relative wear rating 461
resilience effects 455
resistance to wear 163
scrubbing 690
severity effects 467
side slip 457
sliding at zero slip 469
slip dependence 458
stiffness limits 461
temperature dependence 462
tread 673
compound effect 461
Cough model 164
uneven wear 467, 472
wetness effects 466

930
Wet friction 314
Wet road
braking, cornering combination 697
braking force measurement 606
cornering 677, 698
self aligning torque 420
Wet traction
carcass construction effects 343
complexity of problem 332
design for 343
factors affecting 322
mathmatical models for 342
pre-contact problem 348
rib size relation 343
thick films and microtexture 344
tread pattern effects 341
Wettability of rubber 345, 351
Wheel non-uniformity 627
White sidewalls 875
Wholly aromatic polyamide adhesive treatment 107
Wind
friction 584
side force 660
Wobble 629
Worn tires, rolling resistance 594
Wrapping, ply 212

X
X-ray
cord load measurement 570
use in tire inspection 639

Y
Yaw
analysis 785
camber, theoretical relations 786
correction 718
slip relation 304
tangential stress relation 297
velocity equations 788
Young's modulus, rubber 126, 386,
877

z
Zone concept 257

931
* U.S. GOVERNMENT PRINTING OFFICE : 1981 0 - 308-639
V'-" """•' ^' V'1 ' "A*

/ ' 13 9015 00641 7417

Vous aimerez peut-être aussi