Failure Theory (Material)
Failure Theory (Material)
Failure theory is the science of predicting the conditions under which solid materials lose their
strength under the action of external loads. The failure of a material is usually classified into
brittle failure (fracture) or ductile failure (yield). Depending on the conditions (such as
temperature, state of stress, loading rate) most materials can fail in a brittle or ductile manner or
both. However, for most practical situations, a material may be classified as either brittle or
ductile. Though failure theory has been in development for over 200 years, its level of
acceptability is yet to reach that of continuum mechanics.
In mathematical terms, failure theory is expressed in the form of various failure criteria which
are valid for specific materials. Failure criteria are functions in stress or strain space which
separate "failed" states from "unfiled" states. A precise physical definition of a "failed" state is
not easily quantified and several working definitions are in use in the engineering community.
Quite often, phenomenological failure criteria of the same form are used to predict brittle failure
and ductile yield.
Material failure
In materials science, material failure is the loss of load carrying capacity of a material unit. This
definition per se introduces the fact that material failure can be examined in different scales,
from microscopic, to macroscopic. In structural problems, where the structural response may be
beyond the initiation of nonlinear material behavior, material failure is of profound importance
for the determination of the integrity of the structure. On the other hand, due to the lack of
globally accepted fracture criteria, the determination of the structure's damage, due to material
failure, is still under intensive research.
Microscopic failure
Microscopic material failure is defined in terms of crack propagation and initiation. Such
methodologies are useful for gaining insight in the cracking of specimens and simple structures
under well defined global load distributions. Microscopic failure considers the initiation and
propagation of a crack. Failure criteria in this case are related to microscopic fracture. Some of
the most popular failure models in this area are the micromechanical failure models, which
combine the advantages of continuum mechanics and classical fracture mechanics[1]. Such
models are based on the concept that during plastic deformation, microvoids nucleate and grow
until a local plastic neck or fracture of the intervoid matrix occurs, which causes the coalescence
of neighbouring voids. Such a model, proposed by Gurson and extended by Tvergaard and
Needleman, is known as GTN. Another approach, proposed by Rousselier, is based on
continuum damage mechanics (CDM) and thermodynamics. Both models form a modification of
the von Mises yield potential by introducing a scalar damage quantity, which represents the void
volume fraction of cavities, the porosity f.
Macroscopic failure
Macroscopic material failure is defined in terms of load carrying capacity or energy storage
capacity, equivalently. Li[2] presents a classification of macroscopic failure criteria in four
categories:
Five general levels are considered, at which the meaning of deformation and failure is interpreted
differently: the structural element scale, the macroscopic scale where macroscopic stress and
strain are defined, the mesoscale which is represented by a typical void, the micro scale and the
atomic scale. The material behavior at one level is considered as a collective of its behavior at a
sublevel. An efficient deformation and failure model should be consistent at every level.
The failure criteria that were developed for brittle solids were the maximum stress/strain criteria.
The maximum stress criterion assumes that a material fails when the maximum principal stress
σ1 in a material element exceeds the uniaxial tensile strength of the material. Alternatively, the
material will fail if the minimum principal stress σ3 is less than the uniaxial compressive strength
of the material. If the uniaxial tensile strength of the material is σt and the uniaxial compressive
strength is σc, then the safe region for the material is assumed to be
Note that the convention that tension is positive has been used in the above expression.
The maximum strain criterion has a similar form except that the principal strains are compared
with experimentall determined uniaxial strains at failure, i.e.,
The maximum principal stress and strain criteria continue to be widely used in spite of severe
shortcomings.
Numerous other phenomenological failure criteria can be found in the engineering literature. The
degree of success of these criteria in predicting failure has been limited. For brittle materials,
some popular failure criteria are
The approach taken in linear elastic fracture mechanics is to estimate the amount of energy
needed to grow a preexisting crack in a brittle material. The earliest fracture mechanics approach
for unstable crack growth is Griffiths' theory [3]. When applied to the mode I opening of a crack,
Griffiths' theory predicts that the critical stress (σ) needed to propagate the crack is given by
where E is the Young's modulus of the material, γ is the surface energy per unit area of the crack,
and 2a is the crack length. The quantity is postulated as a material parameter called the
'fracture toughness. The mode I fracture toughness is defined as
and is determined experimentally. Similar quantities KIIc and KIIIc can be determined for mode II
and model III loading conditions.
The state of stress around cracks of various shapes can be expressed in terms of their stress
intensity factors. Linear elastic fracture mechanics predicts that a crack will extend when the
stress intensity factor at the crack tip is greater than the fracture toughness of the material.
Therefore the critical applied stress can also be determined once the stress intensity factor at a
crack tip is known.
Energy-based methods
The linear elastic fracture mechanics method is difficult to apply for anisotropic materials (such
as composites) or for situations where the loading or the geometry are complex. The strain
energy release rate[citation needed] approach has proved quite useful for such situations. The strain
energy release rate for a mode I crack which runs through the thickness of a plate is defined as
Where P is the applied load, t is the thickness of the plate, u is the displacement at the point of
application of the load due to crack growth, and 2a is the length of the crack. The crack is
expected to propagate when the strain energy release rate exceeds a critical value GIc - called the
critical strain energy release rate.
The fracture toughness and the critical stain energy release rate are related by
Where E is the Young's modulus. If an initial crack size is known, then a critical stress can be
determined using the strain energy release rate criterion.
Criteria used to predict the failure of ductile materials are usually called yield criteria.
Commonly used failure criteria for ductile materials are:
The yield surface of a ductile material usually changes as the material experiences increased
deformation. Models for the evolution of the yield surface with increasing strain, temperature,
and strain rate are used in conjunction with the above failure criteria for isotropic hardening,
kinematic hardening, and viscoplasticity. Some such models are:
There is another important aspect to ductile materials - the prediction of the ultimate failure
strength of a ductile material. Several models for predicting the ultimate strength have been used
by the engineering community with varying levels of success. For metals, such failure criteria are
usually expressed in terms of a combination of porosity and strain to failure or in terms of a
damage parameter.
Fracture mechanics
Fracture mechanics is the field of mechanics concerned with the study of the propagation of
cracks in materials. It uses methods of analytical solid mechanics to calculate the driving force
on a crack and those of experimental solid mechanics to characterize the material's resistance to
fracture.
Fracture mechanics was developed during World War I by English aeronautical engineer, A. A.
Griffith, to explain the failure of brittle materials.[1] Griffith's work was motivated by two
contradictory facts:
A theory was needed to reconcile these conflicting observations. Also, experiments on glass
fibers that Griffith himself conducted suggested that the fracture stress increases as the fiber
diameter decreases. Hence the uniaxial tensile strength, which had been used extensively to
predict material failure before Griffith, could not be a specimen-independent material property.
Griffith suggested that the low fracture strength observed in experiments, as well as the size-
dependence of strength, was due to the presence of microscopic flaws in the bulk material.
To verify the flaw hypothesis, Griffith introduced an artificial flaw in his experimental
specimens. The artificial flaw was in the form of a surface crack which was much larger than
other flaws in a specimen. The experiments showed that the product of the square root of the
flaw length (a) and the stress at fracture (σf) was nearly constant, which is expressed by the
equation:
An explanation of this relation in terms of linear elasticity theory is problematic. Linear elasticity
theory predicts that stress (and hence the strain) at the tip of a sharp flaw in a linear elastic
material is infinite. To avoid that problem, Griffith developed a thermodynamic approach to
explain the relation that he observed.
The growth of a crack requires the creation of two new surfaces and hence an increase in the
surface energy. Griffith found an expression for the constant C in terms of the surface energy of
the crack by solving the elasticity problem of a finite crack in an elastic plate. Briefly, the
approach was:
Compute the potential energy stored in a perfect specimen under an uniaxial tensile load.
Fix the boundary so that the applied load does no work and then introduce a crack into
the specimen. The crack relaxes the stress and hence reduces the elastic energy near the
crack faces. On the other hand, the crack increases the total surface energy of the
specimen.
Compute the change in the free energy (surface energy − elastic energy) as a function of
the crack length. Failure occurs when the free energy attains a peak value at a critical
crack length, beyond which the free energy decreases by increasing the crack length, i.e.
by causing fracture. Using this procedure, Griffith found that
Where E is the Young's modulus of the material and γ is the surface energy density of the
material. Assuming E = 62 GPa and γ = 1 J/m2 gives excellent agreement of Griffith's predicted
fracture stress with experimental results for glass.
Irwin's modification
Griffith's work was largely ignored by the engineering community until the early 1950s. The
reasons for this appear to be (a) in the actual structural materials the level of energy needed to
cause fracture is orders of magnitude higher than the corresponding surface energy, and (b) in
structural materials there are always some inelastic deformations around the crack front that
would make the assumption of linear elastic medium with infinite stresses at the crack tip highly
unrealistic. F. Erdogan (2000)[2]
Griffith's theory provides excellent agreement with experimental data for brittle materials such as
glass. For ductile materials such as steel, though the relation still holds, the surface
energy (γ) predicted by Griffith's theory is usually unrealistically high. A group working under
G. R. Irwin[3] at the U.S. Naval Research Laboratory (NRL) during World War II realized that
plasticity must play a significant role in the fracture of ductile materials.
In ductile materials (and even in materials that appear to be brittle[4]), a plastic zone develops at
the tip of the crack. As the applied load increases, the plastic zone increases in size until the
crack grows and the material behind the crack tip unloads. The plastic loading and unloading
cycle near the crack tip leads to the dissipation of energy as heat. Hence, a dissipative term has to
be added to the energy balance relation devised by Griffith for brittle materials. In physical
terms, additional energy is needed for crack growth in ductile materials when compared to brittle
materials.
the stored elastic strain energy which is released as a crack grows. This is the
thermodynamic driving force for fracture.
the dissipated energy which includes plastic dissipation and the surface energy (and any
other dissipative forces that may be at work). The dissipated energy provides the
thermodynamic resistance to fracture. Then the total energy dissipated is
G = 2γ + Gp
Where γ is the surface energy and Gp is the plastic dissipation (and dissipation from other
sources) per unit area of crack growth.
For brittle materials such as glass, the surface energy term dominates and
. For ductile materials such as steel, the plastic dissipation term dominates and
. For polymers close to the glass transition temperature, we have
intermediate values of .
Another significant achievement of Irwin and his colleagues was to find a method of calculating
the amount of energy available for fracture in terms of the asymptotic stress and displacement
fields around a crack front in a linear elastic solid.[3] This asymptotic expression for the stress
field around a crack tip is
where σij are the Cauchy stresses, r is the distance from the crack tip, θ is the angle with respect
to the plane of the crack, and fij are functions that are independent of the crack geometry and
loading conditions. Irwin called the quantity K the stress intensity factor. Since the quantity fij is
dimensionless, the stress intensity factor can be expressed in units of .
When a rigid line inclusion is considered, a similar asymptotic expression for the stress fields is
obtained.
Irwin was the first to observe that if the size of the plastic zone around a crack is small compared
to the size of the crack, the energy required to grow the crack will not be critically dependent on
the state of stress at the crack tip.[2] In other words, a purely elastic solution may be used to
calculate the amount of energy available for fracture.
The energy release rate for crack growth or strain energy release rate may then be calculated as
the change in elastic strain energy per unit area of crack growth, i.e.,
Where U is the elastic energy of the system and a is the crack length. Either the load P or the
displacement u can be kept fixed while evaluating the above expressions.
Irwin showed that for a mode I crack (opening mode) the strain energy release rate and the stress
intensity factor are related by:
where E is the Young's modulus, ν is Poisson's ratio, and KI is the stress intensity factor in mode
I. Irwin also showed that the strain energy release rate of a planar crack in a linear elastic body
can be expressed in terms of the mode I, mode II (sliding mode), and mode III (tearing mode)
stress intensity factors for the most general loading conditions.
Next, Irwin adopted the additional assumption that the size and shape of the energy dissipation
zone remains approximately constant during brittle fracture. This assumption suggests that the
energy needed to create a unit fracture surface is a constant that depends only on the material.
This new material property was given the name fracture toughness and designated GIc. Today, it
is the critical stress intensity factor KIc which is accepted as the defining property in linear elastic
fracture mechanics.
Limitations
The S.S. Schenectady split apart by brittle fracture while in harbor (1944)
But a problem arose for the NRL researchers because naval materials, e.g., ship-plate steel, are
not perfectly elastic but undergo significant plastic deformation at the tip of a crack. One basic
assumption in Irwin's linear elastic fracture mechanics is that the size of the plastic zone is small
compared to the crack length. However, this assumption is quite restrictive for certain types of
failure in structural steels though such steels can be prone to brittle fracture, which has led to a
number of catastrophic failures.
Linear-elastic fracture mechanics is of limited practical use for structural steels for another more
practical reason. Fracture toughness testing is very expensive and engineers believe that
sufficient information for selection of steels can be obtained from the simpler and cheaper
Charpy impact test.[citation needed]
Vertical stabilizer, which separated from American Airlines Flight 587, leading to a fatal crash
Most engineering materials show some nonlinear elastic and inelastic behavior under operating
conditions that involve large loads.[citation needed] In such materials the assumptions of linear elastic
fracture mechanics may not hold, that is,
the plastic zone at a crack tip may have a size of the same order of magnitude as the crack
size
the size and shape of the plastic zone may change as the applied load is increased and
also as the crack length increases.
Therefore a more general theory of crack growth is needed for elastic-plastic materials that can
account for:
the local conditions for initial crack growth which include the nucleation, growth, and
coalescence of voids or decohesion at a crack tip.
a global energy balance criterion for further crack growth and unstable fracture.
R-curve
An early attempt in the direction of elastic-plastic fracture mechanics was Irwin's crack
extension resistance curve or R-curve. This curve acknowledges the fact that the resistance to
fracture increases with growing crack size in elastic-plastic materials. The R-curve is a plot of
the total energy dissipation rate as a function of the crack size and can be used to examine the
processes of slow stable crack growth and unstable fracture. However, the R-curve was not
widely used in applications until the early 1970s. The main reasons appear to be that the R-curve
depends on the geometry of the specimen and the crack driving force may be difficult to
calculate.[2]
J-integral
In the mid-1960s James R. Rice (then at Brown University) and G. P. Cherepanov independently
developed a new toughness measure to describe the case where there is sufficient crack-tip
deformation that the part no longer obeys the linear-elastic approximation. Rice's analysis, which
assumes non-linear elastic (or monotonic deformation-theory plastic) deformation ahead of the
crack tip, is designated the J integral.[5] This analysis is limited to situations where plastic
deformation at the crack tip does not extend to the furthest edge of the loaded part. It also
demands that the assumed non-linear elastic behavior of the material is a reasonable
approximation in shape and magnitude to the real material's load response. The elastic-plastic
failure parameter is designated JIc and is conventionally converted to KIc using Equation (3.1) of
the Appendix to this article. Also note that the J integral approach reduces to the Griffith theory
for linear-elastic behavior.
Engineering applications
The following information is needed for a fracture mechanics prediction of failure:
Applied load
Residual stress
Size and shape of the part
Size, shape, location, and orientation of the crack
Usually not all of this information is available and conservative assumptions have to be made.
Short summary
Arising from the manufacturing process, interior and surface flaws are found in all metal
structures. Not all such flaws are unstable under service conditions. Fracture mechanics is the
analysis of flaws to discover those that are safe (that is, do not grow) and those that are liable to
propagate as cracks and so cause failure of the flawed structure. Ensuring safe operation of
structure despite these inherent flaws is achieved through damage tolerance analysis. Fracture
mechanics as a subject for critical study has barely been around for a century and thus is
relatively new. There is a high demand for engineers with fracture mechanics expertise—
particularly in this day and age where engineering failure is considered 'shocking' amongst the
general public.
For the simple case of a thin rectangular plate with a crack perpendicular to the load Griffith’s
theory becomes:
(1.1)
Where G is the strain energy release rate, σ is the applied stress, a is half the crack length, and E
is the Young’s modulus. The strain energy release rate can otherwise be understood as: the rate
at which energy is absorbed by growth of the crack.
However, we also have that:
(1.2)
If G ≥ Gc, this is the criterion for which the crack will begin to propagate.
Irwin's modifications
Eventually a modification of Griffith’s solids theory emerged from this work; a term called stress
intensity replaced strain energy release rate and a term called fracture toughness replaced surface
weakness energy. Both of these terms are simply related to the energy terms that Griffith used:
(2.1)
And
Where KI is the stress intensity, Kc the fracture toughness, and ν is Poisson’s ratio. It is important
to recognize the fact that fracture parameter Kc has different values when measured under plane
stress and plane strain
Fracture occurs when . For the special case of plane strain deformation, Kc becomes
KIc and is considered a material property. The subscript I arises because of the different ways of
loading a material to enable a crack to propagate. It refers to so-called "mode I" loading as
opposed to mode II or III:
Mode I crack – Opening mode (a tensile stress normal to the plane of the crack)
Mode II crack – Sliding mode (a shear stress acting parallel to the plane of the crack and
perpendicular to the crack front)
Mode III crack – Tearing mode (a shear stress acting parallel to the plane of the crack
and parallel to the crack front)
We must note that the expression for KI in equation 2.1 will be different for geometries other
than the center-cracked infinite plate, as discussed in the article on stress intensity. Consequently,
it is necessary to introduce a dimensionless correction factor, Y, in order to characterize the
geometry. We thus have:
(2.4)
Where Y is a function of the crack length and width of sheet given by:
(2.5)
(2.6)
Since engineers became accustomed to using KIc to characterize fracture toughness, a relation has
been used to reduce JIc to it:
The remainder of the mathematics employed in this approach is interesting, but is probably better
summarized in external pages due to its complex nature.
See also
AFGROW - Fracture mechanics and fatigue crack growth analysis software
Fracture toughness
Fatigue
Peridynamics (a numerical method to solve fracture mechanics problems)
Shock (mechanics)
Strength of glass
Strength of materials
Stress corrosion cracking
Stress intensity factor
Strain energy release rate
Structural fracture mechanics
I. PURPOSE AND CONDITIONS
Well constructed failure theories can discriminate safe states of stress in materials from states of
certain failure, based upon calibration by a minimal number of failure type mechanical
properties. The specific purpose here is to provide failure criteria for general types of
materials Two of the conditions that are taken to apply are those of a macroscopic scale of
consideration as well as the corresponding macroscopic homogeneity of the material.
The concepts of macroscopic scale and macroscopic homogeneity have connotations familiar to
everyone. However, trying to define these concepts in absolute terms is extremely difficult.
Macroscopic homogeneity is taken to be the condition that the materials constitution is the same
at all locations. Thus the problem is shifted to the precise meaning of the term location, which
depends upon the scale of observation. Suffice to say, the scale of observation is taken such that
all the common forms of materials are included, such as metals, polymers, ceramics, glasses and
some geological materials. Materials which are excluded are porous materials, whether cellular
or not, as well as granular materials.
In the case of metals, the macroscopic scale must be taken as being much larger than the size of
the individual grains, presumably about an order of magnitude larger. Thus the macroscopic
scale depends upon the type of material. For polymers, it must be much larger than the
characteristic molecular dimension. Other cases are similarly apparent. Effects at the nano-
meter and micro-meter scales do not explicitly enter the macroscopic formulation. To attempt to
bring in these effects would add parameters restricted to a particular materials type and obscure
the clarity of a self contained and self consistent treatment at the macroscopic scale. However,
after obtaining the general macroscopic treatment it is very interesting to compare and interpret
with effects at the smaller scales for specific materials types.
Failure itself can sometimes be difficult to identify and even more difficult to define. For cases
considered here, failure is taken to involve the interruption of the usual linear, reversible range of
behavior by a major change to irreversibility. Failure implies the materials lack of ability to
sustain and support significant loads. In subsequent sections the term damage will be defined as
a subset of failure.
Failure in solids bears an interesting relationship to turbulence in fluids. Fully turbulent flow in
fluids and the failure occurrence in solids mark the departure from linear control to a completely
nonlinear behavior. It does not appear possible to establish an analogy between turbulent flow
and failure in generalized continua; nevertheless both cases conform to the dominance by an
ultimate nonlinearity. In effect both cases represent completion of the account of behavior
prescribed by the balance laws and constitutive equations.
Some perspective may be appropriate on the effort to theoretically characterize the failure of
materials of various types through mathematical criteria. Perhaps no other scientific quest had as
much energy expended upon it historically with so little to show for the effort. The viewpoint
here holds that the scientific advances of the modern era lays out the proper tools and directions
which can finally bring failure criteria to a level of unification (and turbulence modeling as
well). The insight offered by fracture mechanics is especially helpful in understanding brittle
failure in homogeneous materials.
Several qualifications must be included concerning the purpose and scope of this website. It is
not aimed as a literature survey; the goal here is to be technically discriminating rather than
inclusive. The field of materials failure in general and failure criteria in particular is enormously
broad. A complete treatment of such matters would be unrealistic and probably impossible. The
focus here here is upon the theoretical basis for the specific failure criteria under consideration.
By the term “theoretical basis” is meant the physical and mathematical underpinnings of the
various failure forms. No attention will be given to empirical forms which may seem to work
well in certain situations but have no foundation from which to judge their likely generality.
Although it is not intended to integrate specific data sets for particular materials, which itself is a
large and important but self contained topic, nevertheless guides and references to such sources
will be given wherever possible.
Finally, all failure criteria to be reviewed here are taken directly from (or intimately related to)
results already presented in peer reviewed, archive journals. Such publications have received
critical examination and evaluation. While this certainly does not provide any kind of warranty,
it does help to assure fidelity with important aspects of physical behavior, which is the central
objective here. With this emphasis on substance and reliability, the present web based account is
intended as a resource for all those having commitment to materials science and materials
applications.
The next section will be on failure criteria for homogeneous and isotropic materials. Other
sections will be added later starting with the case of highly anisotropic fiber composite
materials. Comments are welcome on anything covered herein.
Richard M. Christensen
August 1, 2007
Fatigue Failure Theories
1. Linear Elastic Fracture Mechanics
2. Strain Life Analysis
3. Stress Life Analysis
Linear Elastic Fracture Mechanics
Assumptions
A crack exist in the part
Around the crack tip, the plastic zone is very small (negligibly small)
When to use it:
When load cycle is less than 1000 cycles
When brittle fracture is most likely mode of failure
Brittle fracture can occur in ductile parts!
How does it work?
The fracture toughness of a material is compared to the stress intensity factor
due to a crack, and if the fracture toughness is higher, the part will not fail.
K
K
fos c
Kc, the fracture toughness for many materials has been determined empirically
and many values for Kc are published in design handbooks, material handbooks,
vendor catalogs, etc.
K must be determined by the engineer. In order to determine K, the stress
intensity factor, the engineer must know the following:
The dimensions of the part being analyzed (2b = plate width)
The length of any existing cracks (or approximations of crack length) (2a =
crack length)
The location of the crack relative to the part’s edge
Is the crack an edge crack?
Is the crack’s orientation perpendicular to the edge of the plate?
The stress induced on the part at the crack tip
What is the nominal stress acting at the crack root?
Are there stress risers in the vicinity of the crack?
Stress intensity factor for an infinitely wide plate is given by:
2
1
K s pa (center crack)
2
1
K 1.12s pa (edge crack)
2
1
K bs pa (if a is not considered << b)
depends on the part’s geometry and the type of loading
Example:
A 10 mm bolt and a 200 mm nut are used to fasten down a plate that may be
assumed to be infinitely wide (a<<b). A uniform stress of 50 N/mm 2 is creating
tension in the plate and the stress intensity factor due to the 10 mm hole in the
plate, Kt, is 3.0. The fracture toughness,Kc, of the steel plate is 40 ksi (in.) 1/2
Will the plate fail in brittle fracture?
What is the critical length that the crack can reach before brittle fracture?
Strain Life Analysis
Assumptions
A crack exist in the part and is growing
When to use it:
When load cycle is less than 1000 cycles
When brittle fracture is most likely mode of failure
Brittle fracture can occur in ductile parts!
How does it work?
Strain data are needed in order to use strain life analysis.
The engineer needs to know total strain and elastic strain:
total elastic e e e
The engineer also needs to know the initial cross-sectional area of the machine
part, Ao, and the final cross-sectional area at the part’s failure.
C=
A
Aln o
The number of cycles to the part’s failure can now be calculated using this
formula:
N 2e C
1
Example:
A strip of mild steel, 10 mm thick, is bent successively about a cylinder of radius
100 mm. It is assumed that for all intents and purposes, the elastic strain may be
safely disregarded. At failure the cross-sectional area of the strip had been
reduced by 60%. How many cycles of bending before the strip will fail due to
brittle fracture?
Stress Life Analysis:
Assumptions:
Cracks do not exist in the part
When to use it:
When load cycle is more than 1000 cycles
When brittle fracture is not catastrophic (deadly!)
What the engineer needs to know:
Part geometry and loading
Part material properties and processing