Fatigue Failure: Design Basis Loads and Qualification

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Fatigue Failure

Fatigue failures start at the most vulnerable point in a dynamically stressed area,
typically a stress raiser, which may be mechanical, metallurgical, or a combination
of two [3].

From: Handbook of Materials Failure Analysis with Case Studies from the Aerospace
and Automotive Industries, 2016

Related terms:

Fatigue Life, Fatigue Crack, Corrosion, Hydrogen, High Cycle Fatigue, Cyclic Load-
ing, S-N Curve, Stress Amplitude

View all Topics

Design Basis Loads and Qualification


George Antaki, Ramiz Gilada, in Nuclear Power Plant Safety and Mechanical Integri-
ty, 2015

What causes fatigue failure?


Fatigue failure is the formation and propagation of cracks due to a repetitive or cyclic
load. Most fatigue failures are caused by cyclic loads significantly below the loads
that would result in yielding of the material. The failure occurs due to the cyclic
nature of the load which causes microscopic material imperfections (flaws) to grow
into a macroscopic crack (initiation phase). The crack can then propagate to a critical
size that results in structural or pressure boundary failure of the component.

Fatigue cracks normally initiate at stress concentrations, structural discontinuities.


Fatigue cracks can also propagate from existing macroscopic cracks, such as weld
defects. Fatigue cracking can be superimposed to a corrosion mechanism, and
the combination of both effects (stress and corrosion) constitutes stress corrosion
cracking (SCC).

> Read full chapter


Repair, Strengthening, and Replace-
ment
Weiwei Lin, Teruhiko Yoda, in Bridge Engineering, 2017

14.3.1.2 Fatigue
Fatigue failure of steel bridges is another significant problem affecting the re-
maining service life of existing steel bridges. In general, fatigue can be defined as
the weakening of steel materials or accumulation of damage at a localized region
caused by cyclic loading or repeatedly applied loads. On this occasion, the material
may damage when the nominal maximum stress is still much less than the material
strength determined from the material tests. When the material is subjected to the
repeated loading above a certain threshold, microscopic cracks will begin to occur
at locations in stress concentration. Then the crack will propagate suddenly causing
the fracture of the steel members.

> Read full chapter

Fatigue failure analysis of welded struc-


tures
Seyed B. Behravesh, ... Hamid Jahed, in Handbook of Materials Failure Analysis with
Case Studies from the Chemicals, Concrete and Power Industries, 2016

Abstract
Fatigue failure analysis is essential for designing welded components subjected to
cyclic loading. Real-world engineering structures will normally have welded com-
ponents, which require a special consideration in the fatigue analysis process. In
this chapter, global and local fatigue modeling approaches for welded structures are
introduced. Local approaches are then discussed in more details and are used later to
estimate the fatigue life of an automotive substructure. The required elements and
the procedure for fatigue failure analysis and life estimation of a welded structure
are explained.

A case study for assessing the fatigue life of an automotive welded substructure is
reviewed. The structure consists of three components, which are made of three dif-
ferent magnesium alloys. The components are joined together using the self-pierce
riveting technique. Monotonic and fatigue test results for the base metals and
welded specimens (with the same joining process as the structure) were obtained
from the literature. The experimental results of the base metals show asymmetric
hardening behavior under tension and compression. A local strain energy method
was employed for fatigue modeling of the base metals and the welded specimens.
To obtain the local strain energy, an asymmetric material model was employed
to account for the asymmetric cyclic hardening behavior of the materials. Fatigue
modeling of the welded structure was performed using the parameters obtained
from the fatigue modeling of the specimens. Finally, the fatigue life and the failure
location predicted by the fatigue modeling showed a good agreement with those
obtained from fatigue testing of the structure.

> Read full chapter

Hydrodynamic Bearings
Luiz Otávio Amaral Affonso, in Machinery Failure Analysis Handbook, 2006

11.4 Fatigue Failures


Fatigue failure of Babbitt metal is similar to any other metal fatigue. The crack
initiation point can be stress concentration caused by an embedded particle, a
high-stress region due to shaft misalignment, high temperature, and so forth. The
fatigue crack may seem to open in the direction of rotation. Pieces of metal may be
carried in the direction of the shaft rotation.

Many fatigue cracks initiate on the interface between Babbitt metal and steel. The
huge difference in elasticity modulus and mechanical strength creates a so-called
metallurgical notch, which facilitates crack initiation in the same way as a mechanical
notch.

Figure 11.3 shows the surface of a bearing that suffered fatigue due to local overload
caused by shaft misalignment. The metal seems to have been peeled from the steel
backing.
FIGURE 11.3. Surface fatigue on a bearing.

> Read full chapter

Fatigue failure of metallic biomaterials


M. Niinomi, in Metals for Biomedical Devices, 2010

Abstract:
Fatigue failure is a significant problem in the usage of metallic biomaterials-
. Understanding the fatigue characteristics of metallic biomaterials is important
for the long term usage of the implants made of metallic biomaterials. Fatigue
strength including fatigue strength level, fatigue strength in vitro and in vivo, notch
fatigue strength, fatigue strength and surface modification, fatigue strength and
surface hardening treatment, fatigue strength and bioactive surface modification,
improvement of fatigue strength including fatigue strength and heat treatment,
fatigue strength and aging, fatigue strength and thermomechanical treatment, fa-
tigue strength and unstable phase, fatigue strength and thermochemical treatment,
fatigue crack propagation characteristics including short and long fatigue crack
propagation in air and in vitro, improvement of long fatigue crack propagation
resistance, fretting fatigue strength in air and in vitro of metallic biomaterials, and
fatigue strength of wire made of new biomedical titanium alloy are described in this
chapter.

> Read full chapter


GENERIC SOLUTIONS FOR REDUC-
ING THE LIKELIHOOD OF OVER-
STRESS AND WEAROUT FAILURES
M.T. Todinov, in Risk-Based Reliability Analysis and Generic Principles for Risk
Reduction, 2007

13.7 GENERIC METHODS FOR REDUCING WEAROUT FAIL-


URES
Fatigue failures are often associated with components experiencing cyclic stresses
or strains resulting in permanent damage. This accumulates until it develops into
a crack which propagates and causes failure. The process of damage accumulation
and failure caused by cyclic loading is called fatigue. Fatigue failures can be reduced
significantly if the development of a fatigue crack is delayed by introducing
compressive residual stresses at the surface. Such compressive stresses delay the
fatigue crack initiation and by causing crack closure also decrease the rate of
crack propagation. One of the reasons is that the compressive stresses subtract
from the loading stresses thereby producing smaller effective stresses. Eliminating
low-strength surfaces can significantly reduce early-life failures due to rapid fatigue
or wear. This can be achieved by:

• eliminating soft decarburised surface after austenitisation;

• eliminating surface discontinuities, folds and pores;

• eliminating coarse microstructure at the surface;

• strengthening the surface layers by surface hardening, carburising, nitriding


and deposition of hard coatings. For example, TiC, TiN and Al2O3 coatings
delay substantially tool wear. Failures due to rapid wear can substantially be
reduced by specifying appropriate lubricants. These create interfacial incom-
pressible films that keep the surfaces from contacting.

Delaying the onset of fatigue failure for hot-coiled compression springs for example
requires (Todinov, 2000b): (i) a small susceptibility to surface decarburisation; (ii)
improved quenching to remove the tensile residual stresses at the spring surface;
(iii) improved tempering to achieve optimal hardness corresponding to a maximum
fatigue resistance; (iv) special surface treatment (e.g. shot peening) resulting in
compressive residual stresses at the spring surface; (v) selecting a cleaner steel with
small number of oxide inclusions which serve as ready fatigue crack initiation sites
and (vi) smaller number density of sulphide inclusions, which promote anisotropy
and reduce the spring wire toughness.
These measures increase the number of cycles needed for fatigue crack initiation,
slow down the rate of fatigue crack propagation and delay significantly the onset
of fatigue failure.

13.7.1 Increasing Fatigue Resistance by Limiting the Size of the Flaws


Usually, the fatigue life of machine components is a sum of a fatigue crack initia-
tion life and life for fatigue crack propagation. The fatigue life of cast aluminium
components and powder metallurgy alloys for example is strongly dependent on
the initiating defects (e.g. pores, pits, cavities, inclusions, oxide films, etc.).

The crack growth rate da/dN is commonly estimated from the Paris–Erdogan power
law (Hertzberg, 1996):

(13.8)

From equation (13.8), the fatigue life N (number of loading cycles) can be estimated
from the integral:

(13.9)

where C and m are material constants and ΔK is the stress-intensity factor range. The
integration limits aiand afare the initial defect size and the final fatigue crack length.
Most of the loading cycles are expended on the early stage of crack extension when
the crack is small. During the late stages of fatigue crack propagation, a relatively
small number of cycles is sufficient to extend the crack until failure. This is the
reason why fatigue life is so sensitive to the size aiof the flaws. Consequently, limiting
the size of the flaws in the material increases the fatigue life.

It has been reported (Ting and Lawrence, 1993) that in cast aluminium alloys, the
dominant fatigue cracks (the cracks which caused fatigue failure) initiated from
near-surface casting pores in polished specimens or from cast surface texture dis-
continuities in as-cast specimens. During fatigue life predictions for such alloys,
the distribution of the initial lengths of the fatigue cracks is commonly assumed
to be the size distribution of the surface (subsurface) discontinuities and pores.
It is also implicitly assumed that the probability of fatigue crack initiation on a
particular defect is equal to the probability of its existence in the stressed volume.
This assumption however is too conservative, because it does not account for the
circumstance that not all pores/defects will initiate fatigue cracks. Fatigue crack
initiation is associated with certain probability because it depends on the type of the
defect, its size and orientation, the mechanical properties and the microstructure
of the matrix, the residual stresses and the stress state in the vicinity of the defect.
Further discussion related to these points is provided in Chapter 16.
13.7.2 Increasing Fatigue Life by Avoiding Stress Concentrators
Sharp notches in components result in high-stress concentration which reduces
fatigue resistance and promotes early-fatigue failures. Such are the sharp corners,
keyways, holes, abrupt changes in cross-sections, etc. Fatigue cracks on rotating
shafts often originate on badly machined fillet radii which act as stress intensifiers.
Because of this, they are reliability critical elements (Thompson, 1999) and their
appropriate design and manufacturing quality should be guaranteed. Reducing the
stress magnitude in the vicinity of a fillet can be achieved by increasing its radius. In
this way, the resistance to an overstress failure or fatigue failure is enhanced. In the
vicinity of a notch, the stress is characterised by a sharp gradient. The smaller the
curvature of the notch, the larger the stress magnitude, the lower the resistance to
overstress and fatigue failures.

Fatigue crack initiation is also promoted at the grooves and the microcrevices of
rough surfaces. These can be removed if appropriate treatment (grinding, honing and
polishing) is prescribed.

Suppose that an elliptical through hole exists in a plate with practically infinite size
relative to the hole. If the major axis of the hole is of length 2a and the minor axis is
of length 2b (Fig. 13.9(a)), the stress at the tips of the major axis can be determined
from (Inglis, 1913):

Figure 13.9. (a) Elliptical hole in a plate; (b) stress concentration around a circular
(bolt) hole in a plate and (c) reducing the stress concentration by appropriate design.

(13.10)

The stress concentration factor is described by the ratio kt= ( A/ . With increasing
the eccentricity of the hole, the stress concentration factor increases. For the special
case of a circular hole (e.g. a bolt hole, Fig. 13.9(b)), according to equation (13.10),
the stress concentration factor becomes kt= (1 + 2a/a) = 3. Reducing the maximum
stress in the vicinity of a hole could be achieved by appropriate design as shown in
the example from Fig. 13.9(c) (Orlov, 1988).

Excessive bending of flexible pipes in dynamic applications for example can also be
associated with excessive localised stresses and strains. Bend restrictors and bend
stiffeners (a bend stiffener is shown in Fig. 13.10) are common design measures to
allow some degree of bending and at the same time to restrict excessive bending. The
bend restrictors consist of interlocking rings around the pipe which do not restrict
decreasing the curvature until a particular critical value is reached. Bending beyond
this critical value causes the rings to lock and no further decrease of the curvature is
possible.

Figure 13.10. Bend stiffener.

13.7.3 Improving Fatigue Resistance by Improving the Condition of the


Surface Layers
The fatigue life of components depends strongly on the condition of the surface.
Numerous observations confirmed that fatigue failures usually start from surface
imperfections. One of the reasons is that the surface layers usually carry the largest
stresses. Furthermore, the surface layer is usually saturated with discontinuities
and defects, and is exposed directly to the negative influence of the processing
and working environment. High-strength steels and alloys are particularly sensitive
to surface defects. Some of the surface imperfections are a direct result from the
manufacturing process. The surface roughness for example is a function of many
parameters: geometry of the cutting tool, type of machined material, homogeneity,
cutting speed and feed, rigidity of the fixtures, vibration resistance of the cutting ma-
chine, degree of wearout of the cutting blade, presence of lubricants and coolants,
etc. The machined surface contains a large number of grooves of different depth
and sharpness, causing local stress concentrations and reduced fatigue strength.
The greater the material strength, the more detrimental the effect of these stress
concentrators. Surface roughness is decreased and fatigue life is improved if the
machined materials have a homogeneous microstructure, characterised by a small
grain size. Surface roughness is decreased by using sharp cutting blades, increasing
the cutting speed, applying lubricants and coolants, eliminating vibrations by using
damping devices and fixtures of high rigidity. The size of the surface irregularities
and their direction has a profound effect on fatigue. Consequently, surface rough-
ness from machining can be reduced significantly and fatigue life further increased
by grinding, polishing honing and superfinish.

Strain-hardening operations such as burnishing, rolling and shot peening increase


fatigue life because the strain-hardened surface layers resist the formation and
propagation of fatigue cracks. As a consequence, in strain-hardened components,
the initiation of fatigue cracks occurs at higher stresses and after a greater number
of loading cycles compared to components which have not been strain hardened.
During burnishing for example, the surface roughness is decreased, surface layers
are strain hardened and residual compressive stresses are generated. Burnishing
also raises the fatigue limit at high temperatures (Zahavi and Torbilo, 1996). As
a result, burnishing applied as a finishing operation to shafts, bars, pistons and
cylinders ensures high reliability.

Even insignificant decarburisation of steels with martensitic structure causes a sig-


nificant reduction of their fatigue strength. Decarburisation diminishes the fatigue
resistance of steel components by: diminishing the local fatigue strength due to the
decreased density of the surface layer, increased grain size and diminished fracture
toughness and yield strength (Chernykh 1991; Todinov, 2000b). These factors create
low cycle fatigue conditions for the surface and promote early-fatigue crack initiation
and premature fatigue failure. Consequently, in order to delay the onset of fatigue
failure, during austenitisation of steel components, decarburisation and excessive
grain growth should be avoided.

> Read full chapter

Fatigue threshold in Mode II and Mode


III, ΔKIIth and ΔKIIIth, and small
crack problems
Yukitaka Murakami, in Metal Fatigue (Second Edition), 2019

Abstract
The problems of fatigue crack growth resistance under mixed mode loading will be
exclusively treated in this chapter.

Fatigue failures of Mode II and Mode III are observed mostly in contact loading
machine components such as bearings, gears, rail road rails, and steel making
rolls. The major loading in these components is cyclic compression coupled with
traction loading. Stable fatigue crack propagation in Mode II or Mode is possible
only under compressive loading. As a Mode II crack coupled with Mode III grows,
the crack finally branches by Mode I. This fact must be noted when we compare
the relative values of ΔKIth, ΔKIIth, and ΔKIIIth. Logically, the value of ΔKIth should
be smaller than ΔKIIth and ΔKIIIIth. To understand the phenomenon related to Mode
II and Mode III fatigue fracture, it is necessary to measure the true values ΔKIIth
and ΔKIIIth without branching to a Mode I crack. This chapter first introduces the
unique method of measurement of ΔKIIth and ΔKIIIth, and afterward the growth
behavior of small cracks in Mode II and III will be discussed based on the precise
observation of the experimental results. The reasons for unsuccessful fatigue life
prediction methods will be explained in the last part of this chapter based on the
observation of complicated crack growth behavior.

> Read full chapter

Fatigue Design of Components


L P Pook, in European Structural Integrity Society, 1997

DISCUSSION
Fatigue failures of components are always a nuisance, and can be expensive and
dangerous. It is therefore not surprising that there are a large number of standards,
written by various organisations, which contain clauses on acceptance fatigue testing
of components[5,6]. Discussion is restricted to documents issued by the British
Standards Institution. It is appreciated that documents issued by other standards
writing organisation, for example the (American) Society of Automotive Engineers,
are also in use in the UK.

It is well known[2] that constant amplitude tests, as compared with realistic variable
amplitude tests, can give a misleading impression of the relative fatigue strengths
of components, especially when the detail design or the material is changed.
Nevertheless, constant amplitude loading is specified in all the acceptance fatigue
tests for components contained in documents listed in Appendix 1. This apparent
contradiction is resolved when commercial considerations are taken into account.
From a commercial viewpoint, an acceptance fatigue test needs to be simple, quick,
and cheap. The risk of a fatigue failure in service may be kept low by making
a simple test relatively severe. In many situations this will minimise the overall
product cost, despite possible overdesign. The 6 tests outlined were chosen to be
representative of those described in British Standards. In all of them the maximum
test duration is relatively short, which minimises both the duration and cost of
testing. It is only in industries where minimum weight design is important, such
as the aircraft industry[1], that expensive and time consuming variable amplitude
tests are economically justified.

As is well known[2], for many metallic materials the results of a fatigue test are largely
independent of the test frequency and waveform. Sinusoidal loading is conventional
in most constant amplitude fatigue testing of specimens and components and,
although not explicitly specified, is used for the acceptance tests on bolts and
car wheels. A square waveform is used for the cookware test for experimental
convenience. The electrical switch test is an operational test, so the waveform is
not specified explicitly. Internal pressure fatigue behaviour is both frequency and
waveform dependent[8]. This is a result of the time required for the pressurizing fluid
to flow into and out of any fatigue cracks that may form. A trapezoidal waveform is
conventionally used for internal pressure fatigue tests for historical reasons[8], and
is therefore specified for the filter housings. Plastics are often strain rate sensitive[9],
an impact fatigue test is therefore specified for the heels of ladies’ shoes as a realistic
representation of service loading.

> Read full chapter

Design and Analysis of Metric Bolted


Joints
Yung-Li Lee, Hsin-Chung Ho, in Metal Fatigue Analysis Handbook, 2012

Fatigue Failure
Fatigue failure occurs when alternating service loading amplitude exceeds the en-
durance limit of a bolt. Fatigue life may reduce significantly if pitting or corrosion
occurs. VDI 2230 adopts the infinite-life design concept for bolted joints by using
the nominal stress-life (S-N) approach. So the following failure criterion applies:

(12.68)

where

AS = fatigue limit = amplitude at 2 × 106 cycles (in N/mm2 = MPa)


SD = safety factor against fatigue (= 1.2 recommended by VDI 2230)
a = nominal stress amplitude, in MPa, calculated on the tensile stress area

Experiments show that fatigue strengths of rolled threads made by rolling after heat
treatment of the bolts, resulting in compressive residual stresses at the thread root,
are better than those rolled before heat treatment. VDI 2230 recommends the fatigue
limit of a bolted joint can be estimated as follows:
• For rolled threads before heat treatment (SV) is(12.69)

• For rolled threads after heat treatment (SG) is(12.70)

where

FSm = mean nominal stress level


F0.2min = minimum yield force = ASRP0.2min

The fatigue strength in the finite life regime where 104cycles ≤ Nf ≤ 2 × 106 cycles
can be calculated

• For rolled threads before heat treatment,(12.71)

• For rolled threads after heat treatment,(12.72)

> Read full chapter

Introduction
D. Radaj, ... W. Fricke, in Fatigue Assessment of Welded Joints by Local Approaches
(Second Edition), 2006

1.1.1 Present state of the art


Fatigue failure of structural members, comprising crack initiation, crack propagation
and final fracture is an extremely localised process in respect of its origin. Therefore,
the local parameters of geometry, loading and material have a major influence on
the fatigue strength and service life of structural members. They must be taken into
account as close to reality as possible when performing fatigue strength assessments
and especially so when optimising the design in respect of fatigue resistance.

Design rules for fatigue-resistant structures, on the other hand, take local effects
only roughly into account. They are based mainly on the nominal stress approach,
which is a global concept in principle. The permissible nominal stresses depend
on the ‘notch class’, ‘detail class’ or ‘fatigue class’ (FAT) of the welded joint being
considered. They are supplemented by general design recommendations.

The code-related state of the art is unsatisfactory in those fields of engineering where
structural members are subjected to fatigue-relevant variable load amplitudes with
appreciable numbers of cycles or where nominal stresses cannot be meaningfully
defined. Local concepts are applied in these areas based on local strain measure-
ments, mainly by strain gauges, and by local stress calculations mainly based on the
finite element method. Both the testing engineer and the structural analyst urgently
need well-founded methods for evaluating these local stresses and strains in respect
of fatigue strength and service life.

These needs can be met only insufficiently if at all. The multitude of proposals on
how to assess the fatigue resistance of structural members based on local parameters
is difficult to overview and evaluate.1 Different fields of engineering, ‘schools’ of
researchers and national communities prefer different approaches. All proposals are
more or less incomplete in respect of user demands, and the local parameter data,
for the most part, lack statistical proof. As a result, the application of local approaches
lags behind the possibilities provided by computerised structural analysis.

> Read full chapter

ScienceDirect is Elsevier’s leading information solution for researchers.


Copyright © 2018 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark of Elsevier B.V. Terms and conditions apply.

You might also like