Sunteți pe pagina 1din 23

FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES

W. V. MARS*
COOPER TIRE & RUBBER COMPANY, 701 LIMA AVE., FINDLAY, OHIO 45840

ABSTRACT
Elastomeric structures subjected to fluctuating loads can fail due to the nucleation and
growth of cracks. To prevent such failures requires understanding of the physics underlying
fatigue failure, approaches for characterizing material behavior, and methods for evaluating the
effects of a given duty cycle. Duty cycles can be complex, often involving simultaneous loading
in multiple directions, and non-periodic variations of the load. By considering how the loads
applied to a structure are transformed into the experiences of individual cracks, a rational frame-
work for predicting fatigue life can be developed.

CONTENTS
II. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .481
II. Fatigue Failure Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .482
III. Physics of Fatigue Failure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .486
A. Crack Driving Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .486
B. Intrinsic Flaws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .486
C. Crack Growth Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .487
D. Crack Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .488
E. Crack Nucleation and Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . .489
IV. Material Behavior . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .490
A. Fully Relaxing Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .491
B. Non-Relaxing Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .492
1. Strain-Crystallizing Rubber . . . . . . . . . . . . . . . . . . . . . . . . . . . .492
2. Non-Strain Crystallizing Rubber . . . . . . . . . . . . . . . . . . . . . . . .494
C. Time-Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .494
1. Tearing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .494
2. Ozone Attack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .495
3. Time-Dependence During Cyclic Loading (the Dwell Effect) .495
D. Approaches to Characterizing Fatigue Behavior . . . . . . . . . . . . . . .496
V. Multiaxial Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .496
A. Scalar Equivalence Theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .496
B. Plane-Specific Equivalence Theories . . . . . . . . . . . . . . . . . . . . . . . .497
VI. Variable Amplitude Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .498
VII. Acknowledgement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .500
VIII. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .500

I. INTRODUCTION
Elastomer structures subjected to fluctuating loads can fail due to the nucleation and growth
of cracks. To prevent such failures, designers need methods for evaluating the many factors
involved in the failure process. Computational methods can be applied early in the product
development cycle, and are therefore particularly attractive, especially considering the costliness
of the traditional cycle of design, build, and break, followed by analysis. Indeed, fatigue testing

* Ph: 419-423-1321; Fax: 419-424-4305; email; wvmars@coopertire.com

481
482 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

may be the most expensive of all product testing: i) it is inherently destructive, ii) it calls for
extended running times, and iii) it requires elaborate systems to apply loading history and col-
lect measurements. There is a need for methods to estimate fatigue life of rubber components
early in the design process.
Much is known about fatigue of rubber. Reviews are available that cover: the physics of
strength and fatigue,1-4 available approaches for fatigue analysis,5 the historical development of
Fracture Mechanics,6 and factors that affect fatigue life.7 Since its founding for rubber,8 over the
last half century, the Fracture Mechanics approach has gained wide recognition, and has matured
into a range of accepted tools that have great power to rationalize crack growth. Understanding
crack growth is a main ingredient in understanding fatigue failure.
Fatigue involves important issues in addition to those addressed by traditional applications
of Fracture Mechanics. These include questions like: How does material response evolve during
a fatigue process? Where will cracks first appear in a structure? What determines the initial ori-
entation of cracks? How do many individual events of different types combine to produce dam-
age and failure?
To reach the aim of fatigue life prediction will require that these and other issues be
addressed by developing capable and mature approaches.9-11 This review aims to provide an
overview of the issues, and to highlight some recent developments.

II. FATIGUE FAILURE PROCESSES


When an elastomer structure experiences a sufficiently long history of fluctuating loads, a
number of changes typically occur. Almost immediately, it will be noticed that the stress-strain
response evolves with each new load cycle. Later, cracks may appear in places where they were
not initially noticed. Cracks will continue growing with each load cycle, often accelerating in
their rate of growth. Eventually, one or another of these processes may prevent the structure from
meeting its purpose.
Figure 1 illustrates typical stress-strain curves collected at several different times during a
fatigue test.11 The solid curve in the figure is the stress-strain behavior observed in a new speci-
men under monotonic increasing load. Other curves in the figure were collected using a second
specimen, during a fatigue test.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 483

FIG. 1. — Uniaxial stress-strain response showing progressive softening due to damage development.

The first few cycles of the test result in significant softening, known as the Mullins effect.12
There is also an associated accumulation of inelastic strain (i.e. during and after the cyclic load-
ing, the unstressed length of the specimen is increased relative to the original unstressed length).
This accumulation of strain is sometimes given the name creep. Creep occurring under the action
of cyclic loading is known to proceed at a much higher rate than predicted from linear vis-
coelastic calculations based on static creep measurements.13,14 The physical origin of viscoelas-
tic creep may be somewhat different than the origin of creep observed under cyclic loading.
Derham and Thomas13 suggest that strain crystallization may be involved in the case of cyclic
loading. In other materials, creep occurring due to cyclic loading is known as ratcheting.15,16
Subsequent cycles result in further softening of the response.17 Figure 1 shows the 8th and
128th cycles, as well as cycles at mid-life and at end-of-life. The rate of softening decreases con-
tinually from its initial high rate to a much reduced rate in mid-life. Figure 2 shows this evolu-
tion, as the stress amplitude decreases with cycles. It is known that the softening rate depends
strongly on the severity of prior events. For example, an overload applied at the beginning of the
history was shown by Mars and Fatemi to nearly eliminate subsequent softening effects.17 These
effects are believed to originate in the molecular structure of the material, as bonds are destroyed
in the elastomer chain network, or between the elastomer and filler particles.
484 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

FIG. 2. — Evolution of stress amplitude with cycles. The middle curve is from the specimen shown in Figure 1.

Late in life, the rate of softening is seen to again increase rapidly. This softening of the
response originates with the rapidly accelerating growth of one or more cracks.18,19 The sequence
of images shown in Figure 3 documents the development of cracks in a uniformly strained test
specimen.11 The length evolution of the largest 8 cracks is plotted in Figure 4. Significant devel-
opment of crack length is seen to occur only late in life, corresponding closely with a rapid loss
in stiffness.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 485

FIG. 3. — Development of cracks in a uniformly strained specimen subjected to cyclic loading.

FIG. 4. — Typical crack length development with cycles.


486 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

III. PHYSICS OF FATIGUE FAILURE


Griffith, in a 1920 paper that helped found the field of Fracture Mechanics,20 and that shaped
the thinking of Rivlin and Thomas,8 wrote: “…the weakness of isotropic solids, as ordinarily met
with, is due to the presence of discontinuities, or flaws, as they may be more correctly called,
whose ruling dimensions are large compared with molecular distances.”
Two decades later, Cadwell and his co-workers21 wrote: “Many theories can be devised to
explain the way the dynamic fatigue properties of rubber vary with the strain and the strain oscil-
lation conditions, but such theories are still conjectures.”
Today we know that Griffith’s hypothesis – that material failure is governed by the evolu-
tion of flaws that were already present in the material - applies not only to the strength of mate-
rials, but also to their fatigue performance. This knowledge provides the basis for understanding
fatigue failure.

A. CRACK DRIVING FORCE

A central question in Fracture Mechanics is how to quantify the generalized force that drives
the growth of a flaw or crack. Rivlin and Thomas8,22 proposed to apply a quantity suggested by
Griffith - the energy release rate (or tearing energy). In this work, the term crack driving force
will be used for this quantity. The crack driving force T is defined as

d (U − V ) (1)
T=−
dA

U is the potential energy associated with various cracked configurations of the material, V is
the work done by external forces as the crack grows. A is the surface area of the crack.
Rivlin and Thomas showed that this parameter unifies fracture8,23 and fatigue crack growth
measurements22,24 from a wide range of test specimens, under a wide range of conditions, and
that it reliably characterizes the loading state of the crack tip. Another critical innovation in their
work was the establishment of test specimens for which Equation 1 can be evaluated analytical-
ly, without having to know details of the stress field. A number of such specimens have emerged,
and are now widely known and applied2 in quantifying the strength and fatigue crack growth
characteristics of rubber. These include the simple tension, pure shear, trousers, and angled test
specimens.
The evaluation of Equation 1 for cracks in “real-world” components cannot generally be
made via a closed form solution. Fortunately, the advent of Finite Element Analysis has enabled
the calculation.25 It has been implemented using various schemes,26-32 and is now widely
applied.33-39

B. INTRINSIC FLAWS

Gent, Lindley, and Thomas40 showed that the fatigue lives of apparently uncracked tension
specimens could be rationalized in terms of the crack growth characteristics of the rubber. They
used the assumption that microscopic flaws occur naturally in the rubber, with a characteristic
effective maximum size (in this case) of 25 x 10-6 m. Their results are shown in Figure 5. The
crack growth characteristics are given in the left-hand plot, showing the crack growth rate dc/dN
as a function of tearing energy T. The fatigue lives N of uncracked specimens, each run at a dif-
ferent strain level and tabulated here as a function of the strain energy density W, are plotted on
the right-hand side.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 487

FIG. 5. — Gent, Lindley, and Thomas40 used the crack growth characteristics (left) to compute the fatigue life of
apparently uncracked dumbbell tension specimens (right). The solid line from the left-hand plot corresponds
to the solid line in the right-hand plot, assuming an initial flaw size of 25 x 10-6 m.

Others applying various methods10,11,18,41-49 have confirmed that effective flaw sizes typical-
ly occur in the range 20 x 10-6 m to 200 x 10-6 m. The effective flaw size depends on many fac-
tors, reflecting that flaws originate from myriad causes. Factors that may impact flaw size
include polymer type,41 crosslink density,41,47 carbon black type,41,49 degree of dispersion,50
mold lubricants11, particulate compound ingredients,51 and microdomain size (controlled by pro-
cessing).52 Often, flaws can be imaged; Figure 6 shows a typical example.11 Le Cam et al.51 also
give images.

FIG. 6. — Fatigue cracks grow from flaws that occur naturally in the rubber.

C. CRACK GROWTH DYNAMICS

Failure due to fracture occurs because the driving force acting on each flaw in the rubber
destroys bonds in the vicinity of flaw boundaries (i.e. the flaw/crack tip), thereby extending flaw
dimensions and surface area. Le Cam et al.51 have recently given detailed evidence on the mech-
anism of this process.
A general approach to modeling the dynamics of the crack growth process has the mathe-
matical form
488 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

dc
dt
(
= f T (t ), T˙ ) (2)

where c is crack length; and t is time, dc/dt is the time rate of crack extension; and T(t) is the time
history of the crack driving force; and Ṫ is the rate of the crack driving force. f is a function (typ-
ically derived empirically) that characterizes the material by relating crack driving force to crack
growth rate.
Equation 2 may be applied to crack growth occurring under either static or dynamic condi-
tions. Fatigue problems, by definition, involve dynamic loading, and are reckoned per cycle N.
Lake and Lindley53 proposed the following specialization, which gives an additive decomposi-
tion of the cyclic crack growth rate r

dc ∂c ∂c dt
=r= + (3)
dN ∂N dynamic ∂t static dN

The partial derivatives in 3 are empirically derived functions giving respectively the cycle-
and time- dependent rates of crack growth. The cycle-dependent part g is a function of the lim-
its of oscillation Tmax and Tmin of the cycle of crack driving force.

∂c
∂N
(
= g Tmax , Tmin ) (4)
dynamic

The time-dependent part h is a function of the instantaneous crack driving force

∂c
= h(T (t )) (5)
∂N static

D. CRACK EVOLUTION

If the crack path, shape, and initial dimensions are known, then Equation 3 contains enough
information to describe the entire evolution of the crack, from its initial, naturally occurring state,
to its final, failure-causing state. Unfortunately, computing the path and shape requires addition-
al assumptions about how crack growth direction is determined, and it remains a challenging task
algorithmically and computationally.
An important special case to consider, for which Equation 3 can be readily solved,54,55 is that
of a crack embedded at the origin of an otherwise uniformly strained, unbounded medium. If the
crack grows in a self-similar manner, in a known plane, then dimensional considerations estab-
lish8,54,56 that the crack driving force T can be written as the product of a nondimensional factor
K, a parameter W̃ quantifying the volume density of energy available for release from the
uncracked medium, and a characteristic length c of the crack.

˜
T = KWc (6)

Expressing the crack driving force as a function of crack size via 6 enables application of
Equation 3 to obtain a 1st order differential equation that can be integrated to obtain the fatigue
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 489

life as the number of cycles required for the crack to grow from initial size c0 to final size cf.

cf 1
N=
∫ c0 r(T (c))
dc (7)

E. CRACK NUCLEATION AND GROWTH

Although the same physical crack growth mechanism governs both stages of the fatigue fail-
ure process, it is useful to idealize the process as consisting of two stages: an initial stage during
which cracks are said to nucleate, and a final stage during which nucleated cracks grow until
structural failure.5,57,58 For example, in Figure 7, crack nucleation is said to occur when a natu-
rally occurring flaw of size c0 develops into a crack of size cf. The number of cycles Nf required
to produce a crack of size cf is the crack nucleation life. The usefulness of idealizing crack
growth as having distinct nucleation and growth phases lies in providing a boundary that gives
guidance on how to best approach the problem.

FIG. 7. — Crack nucleation and growth during a typical fatigue failure.

The nucleation stage involves the potential growth of very many microscopic flaws. Every
material point within a structure is considered as a potential failure location, in order to identify
the most critical location.7,10 The task calls for an efficient method of estimating the driving force
on many cracks under the full range of possible loading conditions. This can be accomplished
via a continuum equivalence parameter.59-61 Fatigue crack nucleation life at each location can be
estimated as a function of the continuum equivalence parameter (i.e. the so-called S-N or
Wohler62 curve).
The crack growth stage focuses attention on the development of one particular crack, or pos-
sibly a small number of specific cracks. The rate of growth of the crack is considered, in order
to evaluate its allowable growth. This task calls for a detailed evaluation of the forces driving the
crack front. The crack driving force varies along the crack front, and depends strongly on details
of the geometry of the structure and the crack.
Figure 8 illustrates the typical evolution of crack driving force at a critical flaw, as the flaw
grows from initial size c0, to crack nucleation at cf, and beyond. The defining feature of the nucle-
ation stage is that the crack driving force varies linearly with the crack size, and is independent
of other length scales of the structure.11,56 After nucleation, the crack driving force comes to be
490 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

governed by additional length scales, generally resulting in non-linear dependence of the crack
driving force on crack size. Thus, cf marks the last point at which the crack driving force may be
estimated, via Equation 6, by regarding the flaw as if it were a crack embedded in an infinite
medium.

FIG. 8. — During the nucleation stage, the crack driving force scales linearly with crack size,
and is independent of structure dimensions. During the growth stage, the crack driving
force is coupled to both crack size and to structure dimensions.

Because microscopic flaws grow so slowly during the nucleation stage, relative to the rapid
growth rates that occur after nucleation, the crack nucleation stage often occupies the most sig-
nificant portion of the total fatigue life. In such cases, it may be acceptable to neglect, for pur-
poses of analysis, the relatively small impact of crack growth beyond the nucleation stage.
There are other situations where a distinct crack nucleation stage may not be significant. In
this case, the entire evolution can be treated as a crack growth problem. Such cases occur when
the initial flaw is situated within a very strong stress gradient,54 effectively rendering cf ≤ c0, or
possibly eliminating completely the regime in which Equation 6 may be applied. The singulari-
ty at the edge of a bonded rubber specimen is one example,33,63 the singularity at the ending of
a steel belt in a tire is another.34

IV. MATERIAL BEHAVIOR


In order to make a prediction about the fatigue life of a structure, knowledge is required of
the fatigue failure characteristics of the material comprising the structure. This knowledge can be
obtained in controlled experiments that apply loading history variations to an appropriate test
specimen. During the experiments, observations are collected of the evolution of the loading his-
tory, and/or the development and growth of cracks.
Loading history variations are generally aimed at quantifying the effects of loading cycle
limits. A typical loading cycle P(t) is shown in Figure 9, along with parameters commonly used
to describe the cycle limits (maximum Pmax, alternating Pa, minimum Pmin, mean Pm, R-ratio R
= Pmin/Pmax). The various parameters are interdependent, because fixing the value of any two of
them uniquely determines the values of those remaining. Thus, the effects of cycle limit varia-
tions can be completely quantified by considering only two independent parameters.7
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 491

FIG. 9 — Commonly used parameters for the description of constant-amplitude mechanical loading history.

In this work, the practice of several pioneers24,41,53,64 is followed, taking maximum and min-
imum load (or equivalently, R-ratio) as the variables of interest. The particular choice of which
parameters to use is arbitrary, so long as two are specified. Many fatigue studies, especially of
non-elastomer materials, take the load amplitude as the primary variable, in recognition of the
difference between a static load (which does not cause fatigue failure, even at high, sub-fracture
levels), and a fluctuating load (which can cause fatigue failure, even at low levels). Mars and
Fatemi7 offer several reasons for preferring maximum and minimum load over combinations
involving the load amplitude. First, both maximum and minimum load are defined for a single
instant that maps directly to a particular material configuration. The mapping makes possible to
track the relationship of the material configuration to the state of cracks embedded in the mate-
rial (are they open or closed?), to determine whether strain-crystallization will occur, and to con-
sider the strong dependence of rubber’s stress-strain behavior on the maximum load12 (i.e. the
Mullins effect). In contrast, the amplitude does not lend itself to such a direct mapping, obscur-
ing the mechanics of the situation.

A. FULLY RELAXING CONDITIONS

First consider fatigue behavior under conditions where the material passes through a fully
relaxed state (i.e. stress-free, Tmin = 0, R = 0) between applications of the load. Although many
structures experience applied loading conditions that, at first glance, do not seem to relax fully
(small cyclic shear superimposed on large compression, for example), it is often the case that the
local experience of the crack at the critical location ends up being a fully relaxing condition.9
This situation may occur especially when compression induces crack closure on at least some
potential planes of failure, and when strain-crystallization retards the growth of competing cracks
that experience non-relaxing conditions.
The fully relaxing condition occurs in structures more often than generally recognized, and
is commonly used in testing to observe several fundamental aspects of fatigue behavior. The
behavior can be modeled via a scheme given below which has been proposed by Mars et al.10
This is a slight modification of Lake and Lindley’s original scheme.41 The new scheme uses a
parameter set that avoids the obscure units arising in the original.
When the maximum driving force Tmax remains below a threshold value To, cracks do not
grow due to mechanical fatigue, suggesting the possibility of infinite life.65 Under such condi-
tions, however, other factors may still eventually cause mechanical failure.7,41 These factors are
492 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

generally environmental or thermochemical in nature, and include the growth of cracks under
steady load (due to ozone or chemical attack), changes to the operating load (due to embrittle-
ment of the elastomer), or evolution of the threshold itself.
When To<Tmax<Tt, where Tt is the point of transition to power-law behavior, the crack
growth rate dc/dN increases linearly with Tmax, according to the following relationship, in which
A is a proportionality constant.

dc
dN
(
= A Tmax − T0 ) (8)

When Tt<Tmax<Tc, where Tc is the critical crack driving force for unstable fracture, the crack
growth rate and driving force are related via a power law,22 with slope F. When Tmax> Tc, the rate
of crack growth increases rapidly from a maximum finite value rc to very high speeds.
F
dc ⎛T ⎞
= rc ⎜ max ⎟ (9)
dN ⎝ Tc ⎠

The coefficient A is not an independent parameter, since it can be computed from the
requirement that the crack growth rate should have the same value on both sides of the transition
point Tt.

rc Tt F
A=
( )
(10)
TcF Tt − T0

B. NON-RELAXING CONDITIONS

It often occurs that the load is not fully relieved between repeats of the loading cycle (i.e.
Tmin > 0, R > 0). Such non-relaxing conditions may strongly influence fatigue behavior.21,64,66,67
The effects depend primarily on the degree to which the elastomer exhibits strain crystallization.
If a material strain-crystallizes, non-relaxing conditions can greatly retard crack growth.64
Otherwise, non-relaxing conditions can modestly speed crack growth.53

1. Strain-Crystallizing Rubber. — The influence of strain crystallization is surprisingly


strong, even for conditions that relax “almost” fully. For example, Figure 10 shows Harbour’s
results,68 similar to earlier results published by Lindley,64 for a strain crystallizing material - gum
natural rubber. In this case, when Tmin is just 5% of Tmax (R = 0.05), the rate of crack growth
decreases by a factor of 4, relative to the fully relaxing case! “Sideways” crack growth69 under
such conditions have been reported, and can complicate measurement efforts.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 493

FIG. 10. — Effect of non-relaxing cycles on fatigue crack growth in strain-crystallizing (left)
and non-strain-crystallizing (right) materials.68

Mars and Fatemi70 proposed a phenomenological model to describe the effect of non-relax-
ing cycles for strain-crystallizing rubber in the range 10-8< dc/dN <10-3 mm/cyc. The model is
based on a few observations from several different sets of data.11,21,64 First, that when plotted
against Tmax, crack growth data at a given level of R tend to exhibit power-law behavior. Second,
that power-law curves for different R ratios tend to converge approximately at a single point. The
point can be estimated as the intersection of: i) the asymptote associated with static load (R = 1),
at which the crack growth rate must change from 0 to +∞, and ii) the power-law curve associat-
ed with the fully relaxing test condition. These observations are embodied by regarding the
power law exponent F in Equation 9 as an empirical function F(R).
Two forms for F(R) have been proposed. The first is a third order polynomial in which the
constants F0, F1, F2, and F3 characterize the material’s dependence on R.

F( R) + F0 + F1 R + F2 R 2 + F3 R3 (11)

the second version is simpler, containing only two parameters F0 and F4:

F( R) = F0 e F4 R (12)

Ostoja-Kuczynski et al.71 have proposed and studied an alternative model, summarized


below in a slightly modified format. Their model considers the range of crack driving force ΔT,
and regards that the parameter rc varies with R ratio, i.e. rc = rc(R).
494 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

F
dc ⎛ ΔT ⎞
= rc ( R)⎜ ⎟
(13)
dN ⎝ Tc ⎠

For 0<R<1, the function rc(R) is given by

rc ( R) = rc × 10 ( 1
− b R 2 + b2 R )
(14)

where b1 and b2 are material parameters.

2. Non-Strain-Crystallizing Rubber. — For materials that do not strain crystallize, non-relax-


ing conditions may accelerate crack growth, and it may be important to consider time-dependent
crack growth. Available models for non-strain-crystallizing elastomers resemble models for other
metallic and polymeric materials,72-75 depending primarily on load amplitude. The simplest of
these models is the well-known Paris law.76 For small R ratios, these models predict almost no
effect, in accord with experiments68 such as those depicted in Figure 10.
Aglan, Moet, and Chudnovsky77,78 proposed a model based on thermodynamic arguments
about the “active zone” preceding the crack tip. The model assumes that the energy dissipation
associated with crack growth is proportional to the square of the energy-release rate range, in
analogy to the Dugdale strip-yield model.79,80 In addition to the critical energy release rate Tc,
the model uses material parameters β and μ.

dc βΔT 2
= (15)
dN μTc − ΔT

Chow and Lu72,73 also proposed a model based on thermodynamic arguments. The model
uses material parameters β and m.

dc βΔT m
= (16)
dN Tc − Tmax

C. TIME-DEPENDENCE

Several conditions may cause rates of time-dependent crack growth in elastomers to match
or exceed cycle-dependent rates of growth. In non-strain-crystallizing elastomers, or at loads
near the failure load, viscoelastic creep and crack growth (“Tearing”) may occur, independent of
mechanical loading cycles. Environmental attack may also occur due to ozone or other chemical
agents. Evolution of elastic and fatigue properties with ageing may introduce additional time-
dependence in the long term.81-83 Finally, it has been recently observed84 that time-dependence
can strongly influence fatigue crack growth rates during cyclic loading, under certain conditions.

1. Tearing. — The driving force required to sustain the steady propagation of a crack
depends on the viscoelastic properties of the elastomer.4 Viscoelasticity is considered to arise
from the time- and temperature- dependent mobility of molecular chains within the elastomer
network. The influence of viscoelasticity was demonstrated by Mullins,85 in experiments on
unfilled vulcanizates of butadiene-styrene and butadiene-acrylonitrile elastomer blends. This
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 495

work, and subsequent studies,86-88 established that the relationship between the crack driving
force, the rate of steady tearing, and the temperature follows laws similar to those that govern
other aspects of viscoelastic materials. Higher damping is associated with higher required ener-
gy release rates to drive a crack at constant velocity.

2. Ozone Attack. — Due to the stress concentration, elastomer network chains at a crack tip
are predisposed to react with ozone. Ozone reacts with carbon-carbon double bonds in the main
polymer chain, causing scission of the chain. Other chemical agents can attack rubber in a man-
ner similar to ozone.
Crack growth due to ozone attack occurs whenever the crack driving force exceeds a small
threshold Tz.89-91 Tz is typically much smaller than the mechanical fatigue threshold To. The value
of Tz depends strongly on compound formulation, and particularly on the presence of antioxi-
dants and antiozonants. For unprotected rubber, Tz ≈ 0.1 J/m2. The presence of antiozonants can
increase Tz by a factor of 10 or more.
Above the threshold value Tz, the rate of crack growth dc/dt due to ozone attack is inde-
pendent of the crack driving force. The crack growth rate depends on the ozone concentration
and on the temperature Θ. Two mechanisms control the crack growth rate. At temperatures near
the glass transition Tg, the crack growth rate is proportional to the temperature, independent of
ozone concentration. At sufficiently high temperatures (Θ-Tg > 100 °C), the crack growth rate
depends only on ozone concentration, and is independent of temperature. The total rate of crack
growth can be estimated via the following model due to Gent and McGrath.92
−1
dc ⎛ 1 1⎞
=⎜ + ⎟ (17)
dN ⎝ R1 R2 ⎠

R1 = K z cO (18)
3

(
F Θ − Tg ) v

R2 = Kv e (
Gv + Θ − Tg ) (19)

R1 reflects the rate at which ozone molecules impact elastomer network chains at the crack
tip. R2 is related to viscoelastic behavior, through the finite time required to redistribute crack tip
cm L
strains, after scission of a network chain. Typical values for these parameters are: K = 0.8 × 10 sec
z
−3
mg ,
cm
−12
K = 0.12 × 10
v
sec
, Fv=40, Gv=52 °C . These constants were derived92 for a series of butadiene-styrene
co-polymers with styrene content in the range 0-80%.

3. Time-Dependence During Cyclic Loading (The Dwell Effect). — Harbour et al.84 have
shown that the rate of crack growth in filled SBR can be highly sensitive to the duration of an
unloaded period occurring during an otherwise cyclic, constant amplitude loading history. A sim-
ilar effect was observed in filled NR, but the magnitude was much smaller. The highest rates of
crack growth are associated with longer unloaded periods, and with fewer cycles between
unloaded periods. It is proposed that the effect is caused by time-dependent recovery in the rub-
ber microstructure at the crack tip, which produces a localized and temporary elevated stress-
state during loading events immediately following the unloaded period.
496 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

D. APPROACHES TO CHARACTERIZING FATIGUE BEHAVIOR

Characterizing via a crack growth approach can provide a meaningful and efficient descrip-
tion of a wide range of fatigue phenomena, over a wide range of conditions. Methods are avail-
able that produce, from a single test specimen, results spanning a range of conditions.93-97
Combined with knowledge of the size of critical flaws, a relatively complete picture can be con-
structed of fatigue crack nucleation behavior,40 with minimum effort.
Alternative approaches, empirically more direct, can also be employed to characterize
fatigue crack nucleation behavior. In these approaches, fatigue crack nucleation life is tabulated
directly as a function of the load amplitude and mean (following Haigh’s approach98-101), as a
function of range and minimum (Cadwell et al.21, or Fielding66), or via a similar scheme (Wohler
curves102). Complete results following such approaches are generally more difficult to obtain,
and are consequently less frequently reported. Each failed test specimen yields a data point for
only a single operating condition. Many tests must be conducted to map out a complete picture
of the fatigue behavior.

V. MULTIAXIAL LOADING
Multiaxial loading occurs whenever loading is applied simultaneously from more than one
direction. While tests for characterizing material behavior are generally made under uniaxial
conditions, the typical duty of a structure imposes multiaxial loading. Whether using a Fracture
Mechanics approach, or a crack nucleation approach, it is important to understand how the many
loads applied to a structure at the global scale are experienced by a crack on the local scale. If
applying a Fracture Mechanics approach, then Equation 1 provides a well-established basis1 for
relating global loads to the local crack experience, and to material characterization results. If
applying a crack nucleation approach, a suitable basis can be selected from among several theo-
ries.5,59-61,98,99,103-105 The reader should note that, at this time, available theories for dealing with
crack nucleation under multiaxial loading conditions have not yet matured to the point where
there is wide consensus about which may be the “most correct”. A few rational guidelines for
selecting among theories have been discussed by Mars.11 In particular, Mars proposes that plane-
specific theories are to be preferred over scalar theories, when multiaxial loading is involved.

A. SCALAR EQUIVALENCE THEORIES

Crack nucleation theories quantify loading conditions in terms of parameters that can be
defined for a material point, in the sense of continuum mechanics.5,60 The maximum principal
value of the stress and strain tensors, and the strain energy density are well-known examples of
such parameters.5 Each of these parameters is a scalar – it has a single component value at a
given material point.
Scalar parameters do not refer to a specific material plane.61 For example, the strain energy
density contains no information that indicates the failure plane. Although there is a normal plane
associated with the maximum principal stress or strain, this plane is associated with the loading
condition, and not with the material. This can be appreciated by considering the experience of a
material plane under a loading history in which the maximum principal stress remains constant,
but the direction rotates. Despite the fact that the maximum principal stress remains constant
(falsely implying infinite fatigue life), a given material plane will experience fluctuating loads
(and a finite fatigue life). There are instances (constant stress state directions, tension) where the
maximum principal stress direction happens to coincide with the normal to the cracking plane,
but these are the exception – not the rule.
Among the scalar theories, it should be pointed out the strain energy density provides an
especially poor correlation61 of results across different multiaxial states.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 497

B. PLANE-SPECIFIC EQUIVALENCE THEORIES

It has been demonstrated experimentally,57,106,107 for a wide range of loading histories


involving combinations of axial and shear loading,108 that crack nucleation tends to occur on one
or more specific material planes.9,10 Typical failure planes from these experiments are illustrated
in Figure 11.

FIG. 11. — Failure planes observed in axial / twist fatigue experiments.57 Path A = Tension; Path B = Torsion;
Path D = Proportional Tension / Torsion; Path C = Fully Reversed Torsion; Path E = Fully
Reversed Proportional Tension / Torsion; Path K Fully Reversed Torsion with Static Tension.

The cracking energy density, proposed recently by Mars,59 is defined for a material point.
The cracking energy density is intended to represent that portion of the stored strain energy den-
sity that may be said to be available for release by virtue of crack growth on a specified materi-
al plane.56 The relationship of the cracking energy density at small strains to results computed
from linear elastic fracture mechanics is shown in Figure 12. The cracking energy density pro-
vides an imperfect, but useable, approximation of the relationship over a wide range of orienta-
tions and multiaxial conditions. At a given material point, the cracking energy density is a func-
tion of the orientation of the cracking plane. The cracking plane is specified through its unit nor-
mal vector.

FIG. 12. — For a wide range of biaxial conditions, and crack orientations, the cracking energy density (left) approxi-
mates behavior computed via linear elastic fracture mechanics (right).56
498 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

Plane specific approaches are needed to properly relate a multiaxial loading history to the
experience of the failure plane, especially when rotation of the loading direction is involved.
With a plane specific approach, the failure plane is predicted by considering the damage accu-
mulated on all possible failure planes, and selecting the plane with the most damage.57,109 Mars’
cracking energy density provided an initial example9 of a plane-specific approach. Saintier et
al.99 and Verron et al.105 have also recently proposed parameters that can consider material plane
orientation. In addition to providing a proper accounting of the effects of loading state rotations,
plane specific parameters provide a means to consider the effects of crack closure due to com-
pression.

VI. VARIABLE AMPLITUDE LOADING


The loading history experienced by a crack in a structure during service usually includes a
diverse range of events, each of which contributes damage at a rate dependent on the extreme
load limits of the event. In order to predict the fatigue life, or the number of times the given load-
ing history may be repeated, the damaging effects of all events in a signal must be considered.
Methods for this task have been developed110-112 and are well known, although they have been
very rarely evaluated or discussed in the open rubber literature. One example discusses the appli-
cation of the Rainflow111 procedure and Miner’s rule112 to fatigue analysis of a rubber mount.113
Another discusses application to fatigue test acceleration.114 Sun et al.115 studied the suitability
of Miner’s linear damage rule, which predicts no dependence on the sequence of loading events.
Their experiments detected a sequence effect, which was attributed to nonlinearity of rubber’s
fatigue behavior (deviation from power law) and elastic behavior (i.e. the Mullins effect). Time
dependent effects, such as those observed during the annealing experiments of Roland and
Sobiesky,116 and in the dwell period experiments of Harbour et al,84 can also influence the accu-
racy of a linear damage rule.
Despite the aforementioned exceptions, a linear damage rule can be used to compute the

overall crack growth r occurring due to a signal containing m different types of events, each
repeated Ni times. A crack growth rate ri corresponds to each event type, and is determined from
characterization tests made under constant amplitude conditions.

m
r= ∑N r i i (20)
i =1

Harbour et al.117 studied the applicability of Equation 20 for a series of 11 different variable
amplitude signals, which are shown in Figure 13. The 11 signals were selected to provide varia-
tions of load level, R-ratio, and load sequence. Such variations are typical of many “real-world”
service histories. These experiments showed that Equation 20 predicts overall crack growth rate
for the 11 signals considered, usually to within a factor of 2, and always to within a factor of 3.
The comparison is shown in Figure 14.
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 499

FIG. 13. — Variable amplitude test signals employed by Harbour et al.117 to investigate the effects on fatigue crack
growth of R-ratio, load level, load sequence and dwell period. Signals are plotted as displacement δ vs. time t.

FIG. 14a. — Comparison of crack growth rate predictions117 from Equation 20 with experimentally
observed crack growth rates for variable amplitude test signals 1 through 11, in filled NR.
Broken lines indicate factors of 2 and 3 difference from perfect agreement, respectively.
500 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

FIG. 14b. — Comparison of crack growth rate predictions117 from Equation 20 with experimentally observed crack
growth rates for variable amplitude test signals 1 through 11, in filled SBR.
Broken lines indicate factors of 2 and 3 difference from perfect agreement, respectively.

An additional signal, which included a dwell period, produced higher crack growth rates
than the corresponding constant amplitude test signal containing no dwell period.84 For a filled
SBR, these tests produced an average experimental growth rate up to 30 times greater than the
constant amplitude rate. The dwell time and number of applied cycles between dwell periods
were the most significant influences on crack growth rate. Dwell periods also produced increased
crack growth rates in filled NR, but the effect was less significant.

VII. ACKNOWLEDGEMENT
The author would like to acknowledge with gratitude Ali Fatemi and Ryan Harbour for fruit-
ful collaboration on many of the works cited here. Also, the author is indebted to Alan Muhr and
Erwan Verron for helpful input. Finally, the author thanks the Cooper Tire & Rubber Company
for their support and for permission to publish the work.

VIII. REFERENCES
1G. J. Lake, RUBBER CHEM. TECHNOL. 76, 567 (2003).
2Engineering with Rubber, "How to Design Rubber Components," A. Gent, Ed., published by Carl Hanser Verlag,
Munich, 1992.
3Science and Technology of Rubber, F. R. Eirich, Ed., Academic Press, 1978, pp. 419-454.
4B. N. J. Persson, O. Albohr, G. Heinrich, H. Ueba, J. Phys.: Condens. Matter 17, R1071 (2005).
5W. V. Mars, A. Fatemi, Int. J. Fatigue 24, 949 (2002).
6A. G. Thomas, RUBBER CHEM. TECHNOL. 67, G50 (1994).
7W. V. Mars, A. Fatemi, RUBBER CHEM. TECHNOL. 77, 391 (2004).
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 501

8R. S. Rivlin, A. G. Thomas, J. Polym. Sci. 10, 291 (1953).


9W. V. Mars, A. Fatemi, RUBBER CHEM. TECHNOL. 79, 589 (2006).
10W. V. Mars, J. Kingston, A. Muhr, S. Martin, K. W. Wong, Stockholm Royal Institute of Technology in Fatigue Life
Analysis of an Exhaust Mount; “Constitutive Models for Rubber IV,” Lisse, A.A. Balkema, 2005, p. 23-29.
11W. V. Mars, Ph.D. Dissertation, University of Toledo, 2001.
12L. Mullins, RUBBER CHEM. TECHNOL. 42, 339 (1969).
13C. J. Derham, A. G. Thomas, RUBBER CHEM. TECHNOL. 50, 397 (1977).
14G. B. McKenna, L. J. Zapas, RUBBER CHEM. TECHNOL. 54, 718 (1981).
15T. Hassan, S. Kyriakides, Int. J. Plasticity 8, 91 (1992).
16X. Chen, S. Hui, Polymer Testing 24, 829 (2005).
17W. V. Mars, A. Fatemi, J. Eng. Mat. Technol. 126, 19 (2004).
18W. V. Mars, A. Fatemi, Fatigue Fracture Eng. Mater. Struct. 26, 779 (2003).
19W. V. Mars, A. Fatemi, J. Mater. Sci. 41, 7324 (2006).
20A. A. Griffith, Phil. Trans. R. Soc. London, Ser. A 221, 163 (1921).
21S. M. Cadwell, R. A. Merrill, C. M. Sloman, F. L. Yost, RUBBER CHEM. TECHNOL. 13, 304 (1940).
22A. G. Thomas, J. Polym. Sci. 18, 177 (1955).
23H. W. Greensmith, A. G. Thomas, J. Polym. Sci. 18, 189 (1955).
24A. G. Thomas, J. Polym. Sci. 31, 467 (1958).
25P. B. Lindley, J. Strain Analysis 7, 13 (1972).
26D. M. Parks, Computer Methods Appl. Mech. Eng. 12, 353 (1977).
27C. F. Shih, B. Moran, T. Nakamura, Int. J. Fracture 30, 79 (1986).
28R. Mueller, G.A. Maugin, Computational Mechanics 29, 52 (2002).
29R. Mueller, S. Kolling, D. Gross, Int. J. Numerical Methods Eng. 53, 1557 (2001).
30P. Steinmann, Int. J. Solids Structures 37, 7371 (2000).
31B. Naser, M. Kaliske, R Muller, Computational Mechanics, DOI 10.1007/s00466-007-0159-9 (2007).
32K. N. Shivakumar, P. W. Tan, J. C. Newman Jr., Int. J. Fracture 36, R34 (1988).
33O. H. Yeoh, Plastics, Rubbers, Compos. 30, 389 (2001).
34T. G. Ebbott, Tire Sci. Technol. 24, 220 (1996).
35Y. W. Chang, A. N. Gent, J. Padovan, Int. J. Fracture 60, 363 (1993).
36O. H. Yeoh, Mech. Materials 34, 459 (2002).
37K. N. Morman, T. Y. Pan, RUBBER CHEM. TECHNOL. 61, 503 (1988).
38J. J. C. Busfield, V. Jha, H. Liang, I. C. Papadopoulos, A. G. Thomas, Plast., Rubber Compos. 34, 349 (2005).
39N. Ait Hocine, M. Nait Abdelaziz, A. Imad, Int. J. Fracture 117, 1 (2002).
40A. N. Gent, P. B. Lindley, A. G. Thomas, J. Appl. Polym. Sci. 8, 455 (1964).
41G. J. Lake, P. B. Lindley, J. Appl. Polym. Sci. 9, 1233 (1965).
42I. S. Choi, C. M. Roland, RUBBER CHEM. TECHNOL. 69, 591 (1996).
43K. Mukhopadhyay, D. K. Tripathy, S. K. De, RUBBER CHEM. TECHNOL. 66, 38 (1993).
44F. Abraham, G. Clauss, T. Alshuth, Stockholm Royal Institute of Technology in Testing and Simulation of the Influence
of Glass Spheres on Fatigue Life and Dynamic Crack Propagation of Elastomers; “Constitutive Models for Rubber
IV,” Lisse, A.A. Balkema, 2005, p. 71-76.
45D. G. Young, E. N. Kresge, A. J. Wallace, RUBBER CHEM. TECHNOL. 55, 428 (1982).
46M. Braden, A. N. Gent, J. Appl. Polym. Sci. 3, 100 (1960).
47G. R. Hamed, RUBBER CHEM. TECHNOL. 56, 244 (1983).
48Y. Fukahori, Int. Polym. Sci. Tech. 13, T/37 (1986).
49E. S. Dizon, A. E. Hicks, V. E. Chirico, RUBBER CHEM. TECHNOL. 47, 231 (1974).
502 RUBBER CHEMISTRY AND TECHNOLOGY VOL. 80

50G. S. Fielding-Russell, R. L. Rongone, RUBBER CHEM. TECHNOL. 56, 838 (1983).


51J.-B. Le Cam, B. Huneau, E. Verron, L. Gornet, Macromolecules 37, 5011 (2004).
52M. N. Qureshi, Ph.D. Dissertation, University of Akron, 2002.
53G. J. Lake, P. B. Lindley, J. Appl. Polym. Sci. 8, 707 (1964).
54J. Gough, A H. Muhr, Stockholm, Royal Institute of Technology in Energy Release Rates for Small Cracks in Rubber
Components; “Constitutive Models for Rubber IV,” Lisse, A.A. Balkema, 2005, pp. 51-57.
55R.A. Sack, Proc. Physical Soc. 58, 729 (1946).
56W. V. Mars, in Heuristic Approach for Approximating Energy Release Rates of Small Cracks Under Finite Strain
Multiaxial Lodaing; Covney, V.A. Ed; Elastomers and Components, “Service Life Prediction – Progress and
Challenges,” Woodhead Publishing Limited, Cambridge, 2006, pp. 91-111.
57W. V. Mars, A. Fatemi, J. Mater. Sci. 41, 7324 (2006).
58J. Gough, A. H. Muhr, Proc. Int. Rubber Conference, Maastricht, Publ. IOM Communications, London, 165, (2005).
59W. V. Mars, RUBBER CHEM. TECHNOL. 75, 1 (2002).
60W. V. Mars, A. Fatemi, Fatigue Fracture Eng. Mater. Struct. 28, 515 (2005).
61W. V. Mars, A. Fatemi, RUBBER CHEM. TECHNOL. 80, 169 (2007).
62A. Wöhler, Engineering 2, 160 (1867).
63O. H. Yeoh, RUBBER CHEM. TECHNOL. 76, 483 (2003).
64P. B. Lindley, Int. J. Fracture 9, 449 (1973).
65G. J. Lake, P. B. Lindley, J. Appl. Polym. Sci. 10, 343 (1966).
66J. H. Fielding, Ind. Eng. Chem. 35, 1259 (1943).
67C. Bathias, K. LeGorju, C. Lu, L. Menabeuf, Fatigue and Fracture Mechanics: 27th Volume, ASTM STP 1296, R. S.
Piascik, J.C. Newman, N. E. Dowling, Eds., American Society for Testing and Materials, 1997, p. 505.
68R. Harbour, A. Fatemi, and W. V. Mars, Fatigue Fracture Eng. Mater. Struct. 30, 640 (2007).
69A. Gent, M. Razzaghi-Kashani, G. R. Hamed, RUBBER CHEM. TECHNOL. 76, 122 (2003).
70W. V. Mars, A. Fatemi, RUBBER CHEM. TECHNOL. 76, 1241 (2003).
71E. Ostoja-Kuczynski, P. Charrier, E. Verron, L. Gornet, G. Marckmann, Stockholm, Royal Insitute of Technology in
Influence of Mean Stress and Mean Strain on Fatigue Life of Carbon Black Filled Natural Rubber; “Constitutive
Models for Rubber IV,” Lisse, A.A. Balkema, 2005, p. 15-21.
72C. L. Chow, T. J. Lu, Tire Science Technol. 20, 106 (1992).
73C. L. Chow, T. J. Lu, Int. J. Fracture 40, 53 (1989).
74V. E. Kearny, R. M. Engle, R. G. Forman, J. Basic Engineering 89, 459 (1967).
75J. Weertman, Int. J. Fracture Mechanics 2, 460 (1966).
76P. C. Paris, M. P. Gomez, W. P. Anderson, Trend Eng. 13, 9 (1961).
77H. Aglan, A. Moet, RUBBER CHEM. TECHNOL. 62, 98 (1989).
78A. Chudnovsky, A. Moet, J. Elastomers Plastics 18, 50 (1986).
79D. S. Dugdale, J. Mech. Phys. Solids 8, 100 (1960).
80G. I. Barenblatt, Adv. Appl. Mech. 7, 55 (1962).
81A. Ahagon, M. Kida, M. Kaidou, RUBBER CHEM. TECHNOL. 63, 683 (1990).
82A. R. Azura, A.G. Thomas, in Effect of Heat Ageing on Crosslinking, Scission and Mechanical Properties, V. Coveney,
Ed., Elastomers and Components: Service Life Prediction-Progress and Challenges, Woodhead Publishing,
Cambridge, 2006, pp. 27-38.
83A. R. Azura, A.H. Muhr, D. Goritz, A.G. Thomas, London, U. in Effect of Ageing on the Ability of Natural Rubber to
Strain Crystallise; “Constitutive Models for Rubber III,” Lisse, A.A. Balkema, 2003, pp. 79-84.
84R. Harbour, A. Fatemi, W. V. Mars, RUBBER CHEM. TECHNOL., in press.
85L. Mullins, Trans. Inst. Rubber Ind. 35, 213 (1959).
FATIGUE LIFE PREDICTION FOR ELASTOMERIC STRUCTURES 503

86A. Kadir, A. G. Thomas, RUBBER CHEM. TECHNOL. 54, 15 (1981).


87A. Kadir, A. G. Thomas, J. Polym. Sci. Polym. Phys. Ed. 22, 1623 (1984).
88T. Smith, RUBBER CHEM. TECHNOL. 51, 225 (1978).
89M. Braden, A. N. Gent, RUBBER CHEM. TECHNOL. 35, 200 (1962).
90G. J. Lake, P. B. Lindley, Rubber J. 146 (10), 24 (1964).
91G. J. Lake, P. B. Lindley, Rubber J. 146(11), 30 (1964).
92A. N. Gent, J. E. McGrath, J. Polym. Sci., Part A 3, 1473 (1965).
93A. N. Gent, M. Razzaghi-Kashani, RUBBER CHEM. TECHNOL. 73, 818 (2000).
94H. Aboutorabi, T. G. Ebbott, A. N. Gent, O. H. Yeoh, RUBBER CHEM. TECHNOL. 71, 76 (1998).
95A. J. M. Sumner, S. A. Kelbch, U. G. Eisele, Rubber World 213, 38 (1995).
96D. G. Young, RUBBER CHEM. TECHNOL. 63, 567 (1990).
97D. G. Young, Paper 63 presented at the 168th Rubber Division ACS Meeting - Fall 2005, Pittsburgh PA., November 1-
3, 2005.
98N. Saintier, G. Cailletaud, R. Piques, Int. J. Fatigue 28, 61 (2006).
99N. Saintier, G. Cailletaud, R. Piques, Int. J. Fatigue 28, 530 (2006).
100N. Andre, G. Cailletaud, R. Piques, Kautsch. Gummi Kunstst. 52, 120 (1999).
101T. Steinweger, U. Weltin, M. Flamm, Royal Institute of Technology in Four Tests to Characterize Haigh-Diagram for
Damage Calculations;“Constitutive Models for Rubber IV,” Lisse, A.A. Balkema, 2005, pp. 9-14.
102F. Abraham, T. Alshuth, S. Jerrams, in Parameter Dependence of the Fatigue Life of Elastomers, V.A. Coveney, Ed,
Elastomers and Components: Service Life Prediction - Progress and Challenges, Woodhead Publishing, Cambridge,
2006, pp. 59-73.
103W. V. Mars, A. Fatemi, Fatigue Fracture Eng. Mater. Struct. 28, 523 (2005).
104A. Zine, N. Benseddiq, M. Nait Abdelzaziz, N. Ait Hocine, D. Bouami, Fatigue Fracture Eng. Mater. Struct., 29, 267
(2006).
105E. Verron, Stockholm, Royal Institute of Technology in Prediciton of Fatigue Crack Initiation in Rubber with the Help
of Configurational Mechanics; Constitutive Models for Rubber IV, Lisse, A.A. Balkema, 2005, pp. 3-8.
106W. V. Mars, A. Fatemi, Int. J. Fatigue 28, 521, (2006).
107R. Harbour, A. Fatemi, W. V. Mars, J. Mater. Sci. in press.
108W. V. Mars, A. Fatemi, Exp. Mech. 44, 136 (2004).
109W. V. Mars, U.S. Patent 6,634,236, 2003.
110H. O. Fuchs, R. I. Stephens, “Metal Fatigue in Engineering,” John Wiley & Sons, ISBN 0-471-05264-7, 1980.
111S. D. Downing, D. F. Socie, Int. J. Fatigue 4, 31 (1982).
112M. A. Miner, J. Applied Mech. 6, 159 (1945).
113D. Klenke, A. Beste, Kautsch. Gummi Kunstst. 40, 1067 (1987).
114T. Steinwegger, M. Flamm, U. Weltin, London U. in A Methodology for Test Time Reduction in Rubber Part Testing;
“Constitutive Models for Rubber III,” Lisse, A.A. Balkema, 2003, pp. 27-32.
115C. Sun, A. N. Gent, P. Marteny, Tire Sci. Technol. 28, 196 (2000).
116C. M. Roland, J. W. Sobiesky, RUBBER CHEM. TECHNOL. 62, 683 (1989).
117R. Harbour, A. Fatemi, W. V. Mars, Int. J. Fatigue in press.

[ Based on Paper 5, presented at the Spring Rubber Division, ACS, Meeting, (Akron, Ohio)
April 30-May 2, 2007; revised May 2007 ]

S-ar putea să vă placă și