You are on page 1of 14

Degradation mechanisms essay

Subject: Metal Fatigue

Module: Materials Performance - Life Cycle Design


Page |1

Introduction

According to [1], Fatigue is the progressive, localized, permanent structural change


that occurs when a material is subjected to conditions that produce fluctuating stresses and
strains at some points and that may culminate in cracks or complete fracture after a
sufficient number of fluctuations. The stresses that lead to fatigue are usually, but not
always, significantly lower than the materials yield strength, due to a mechanism involving a
number of stages.

Failure due to fatigue is the most common type of mechanical failure, costing an
enormous amount of money and lives each year. As a result, testing and prediction of
fatigue behaviour is vital during engineering design. [2]

Mechanisms of failure

The Mechanism of failure due to fatigue follows a specific pattern, which according
to the [3] can be distinguished into the following five stages:

Stage 1: Cyclic plastic deformation, prior to crack initiation

This stage precedes the formation of microcracks. The dislocations present in the
material will move and lead to the creation of permanent slip bands and a rough surface,
which can subsequently act as crack initiation sites[4]. This can be seen in diagram /0/:
Under monotonic loading, coarse slip occurs (a), where only a few slip bands move relative
to each other. However, under cyclic loading, fine slip will develop (b) with many slip bands
moving.[2]

At this stage, cyclic hardening (usually in annealed alloys,[5, 6]) or cyclic softening
(usually in cold worked alloys,[6]) can take place.

Stage 2: Initiation of one or more Microcracks

Crack initiation takes place almost entirely on the surface of the material. It takes
place on grain boundaries, second-phase particles or inclusion interfaces[3]. It usually
Page |2

occurs due to the development of the persistent slip bands to cracks according to the
mechanism seen in diagram /0 (c)/[2]: An avalanche of slip movements will eventually lead
to the formation of intrusion-extrusion pairs on the surface (c). These intrusions will develop
into microcracks.

Diagram 0:(a) Steady stress. (b) Cyclic stress. (c) Extrusion/Intrusion pair leading to fatigue

Stage 3: Propagation and coalescence of Microcracks

This stage is also called Stage I of fatigue crack growth. The propagation of the
microcracks takes place along a single crystallographic direction, following a single slip plane
where the shear stress has its maximum value. The propagation speed is exceptionally slow
(a few nm/cycle) and as a result, up to 95% of the total life may be spent in this stage for
high-cycle fatigue [7]. When microcracks reach grain boundaries, some will cease
Page |3

propagating, and others will continue across, following the slip plane of the new grain. This
means that the crack course depends on metallurgical factors.

Stage 4: Propagation of one or more macrocracks

This is also called Stage II of fatigue crack growth. When the crack has propagated a
few grain distances inside the material, it will start propagating faster (μm/cycle). The
propagation will now be perpendicular to the direction of the principal stress.

The surface created during this stage exhibits striations and beachmarks. The first
are microscopic and can only be observed with an electron microscope. Each striation
represents the advance distance of a crack front during a single load cycle, while
beachmarks are visible to the naked eye, and represent interruptions during the crack
propagation stage[8]

Stage 5: Final failure

Also called Stage III of fatigue. The continuous propagation of a Stage II crack results
in the reduction of the area of the material subjected to loading. When the stress on the
remaining area overcomes the material strength, failure occurs almost instantly. A dimple-
like appearance is characteristic of this stage, as rupture occurs by the enlargement and
joining of pores (dimples), causing significant plastic deformation[7]. Brittle fracture, is also
possible, but much less common[8].

Role of microstructure

The microstructure of the material has a big effect on its fatigue behaviour. While it
does not alter the mechanism, it greatly affects crack initiation and Stage I propagation
duration. According to [9], microstructure defects such as inclusions and pores can promote
crack initiation due to the concentration of stresses in the area. Grain boundaries are also
areas where crack initiation is easier, especially in the presence of brittle intermetallics or
impurities. The texture of the material also plays a significant role, as does grain size: larger
grains or adjacent grains with uniform orientation can facilitate crack propagation, as the
crack does not have to change crystallographic orientation[9]. A lamellar structure has also
Page |4

been found to affect fatigue behaviour, with the lamellae acting as barriers to crack
propagation, especially if they are vertical to the stresses[10].

Lifetime prediction

Specimens and testing

The machines used for fatigue testing can subject the specimen in bending, torsion,
axial loading or a combination of the above. The cyclic loading of the specimen is usually
carried out until failure occurs, or a set number of cycles is reached without failure. He data
acquired from these tests are used to plot the S-N curve of the specimen and calculate the
crack growth rates. A number of fatigue testing machines can be seen in diagram /2/.

Diagram 2:(a) Cantilever rotating bending. (b) Rotating pure bending. (c) Axial loading

The specimens used in fatigue testing are distinguished between total fatigue life
testing /Diagram 3 (a) and (b)/ and fatigue crack growth testing, /Diagram 3 (c) and (d)/
Page |5

Diagram 3: Fatigue Testing specimens (a) Rotating bending specimen. (b) Axial or bending
with circumferential groove. (c) Three-point bend (d) Compact tension

The S-N curve

The S-N curve, or Wöhler diagram, was invented by August Wöhler after he
concluded that the fatigue resistance of a material was dependant on the applied cyclic
stress amplitude σα, or S, which is:[7]

σ max −σ min
S = σα =
2

Wöhler conducted fatigue experiments by submitting specimens to symmetrical


cyclic stress (σ max=−σ min) and plotting the number of cycles to failure on a logarithmic scale.
He also noted that, for ferrous alloys, a lower stress amplitude σ f exists, below which fatigue
failure does not happen.

A characteristic S-N curve for a ferrous alloy can be seen in diagram /1/. The curve
can be separated into three distinct regions, according to the type of fatigue that occurs:

1
i. Region I ( - 104 cycles). Here the failure is due to plastic strain, as the loads
4
on the material are close (or above) its yield strength. The material absorbs
energy into its crystal lattice, and as a result a hysteresis loop can be seen in
Page |6

its stress-strain diagram[7]. Fatigue that takes place in this region is called
low-cycle fatigue.
ii. Region II (104 – 107 cycles). Here the behaviour is elastic on a macroscopic
scale, and the load is lower than the material’s yield strength. This area is the
most common in engineering, and is called high-cycle fatigue.
iii. Region III (>107). In this region, failure due to fatigue will not take place, as
the material has reached its fatigue limit. While most ferrous alloys have a
fatigue limit after about 10 6-107 cycles , others, such as aluminium and
copper, do not have one[3]. It should be noted that, while the fatigue limit is
an established and accepted element in fatigue, some newer research
maintains that it does not exists in metals and alloys, and that failure can take
place even after 109 cycles [9, 11]

Diagram 1: S-N curve for a ferrous alloy, with three characteristic regions (high cycle
fatigue, low cycle fatigue, fatigue limit.

Fatigue behaviour

The lifetime of smooth materials can be calculated from their S-N curve. However,
plotting an S-N curve is time-consuming and requires hundreds of experiments for a single
Page |7

component, so a mathematical approximation of the fatigue behaviour is often


necessary[2]. There are different equations addressing low-cycle fatigue and high cycle
fatigue:

Low cycle fatigue

At low cycle fatigue we can use the Coffin-Manson relationship, which describes the
plastic part of the S-N curve:

εα = εf*(N)c

where εα=Δεα/2 is the total strain amplitude, N is the life at ε α, and εf, C coefficients
depending on the material.

High cycle fatigue

At high cycle fatigue, the lifetime of the material can be approximated via Basquin’s
Law, who found that the elastic part of the S-N curve can be described by a power law:[7]:

σα = σf*(N)b

where σα is the stress amplitude, N the life at σα, and σf,b coefficients.

Pre-cracked specimens

When a crack already exists, the fatigue life can be more easily calculated using
Griffith’s theory of brittle fracture. More specifically, for Stage II crack growth, Paris’ Law
can be used(Diagram /10/)[3]:

Da/DN = C*(ΔK)N

Where a is the crack length and N is the number of cycles. ΔK is the stress intensity range,
where ΔΚ = Κmax – Κmin, and K the stress intensity factor.
Page |8

Diagram 10: Paris’ law only applies for Stage II crack growth

Miner’s Rule

Often, a component is subject to cyclic stresses of many different amplitudes and


frequencies, either at the same time or consecutively. When this is the case, fatigue cannot
be predicted immediately using the S-N diagram. Instead, in 1948 Miner proposed that a
linear damage accumulation model that can be used for lifetime prediction[7]:

N
n
∑ Ni =1
i=1 i

In this model, n1 is the number of cycles at stress S i that the material has been
subjected to, in comparison to the maximum lifetime (cycles to failure) N i at this stress.
Failure due to fatigue takes place when the sum of all ni/Ni pairs is 1.
Page |9

Miner’s rule is a useful tool, although it should be used carefully as it has two
significant limitations: Firstly, it does not take into account the loading sequence, and
secondly, it cannot account for the statistical dispersion of the cumulative damage[12].

Diagram 11: Using S-N curve to find Ni for each Si in order to apply Miner’s rule

Addressing fatigue

Criteria for fatigue design

Taking fatigue into account during design is essential when development of a


product. However, depending on the application, there are several different approaches
when designing for fatigue[2].

Infinite life design

Designing for infinite life is the oldest criterion. In order for the designed component
to never fail, the local stresses must be below the fatigue limit. While designing for infinite
life can be useful in a few situations, it is usually expensive and impractical, and has been
replaced by the safe-life design approach[2].
P a g e | 10

Safe-life design

The safe-life design approach takes into account the number of cycles that the
component will undergo during its service life, and is designed so that failure will never
occur during this service life. This design approach demands that components should be
replaced once their service life has been exceeded. It is mainly used in a variety of fields
where regular inspections are not possible[2].

Fail-safe design

The fail-safe approach was developed by the aerospace industry, as safe-life is costly
and wasteful, and service failures cannot be always predicted by testing or analysis. In this
approach, fatigue damage is expected to take place, but this damage should not lead to
failure before it is detected and the part replaced. A component that has been designed
with the fail-safe design must be inspected for cracks at regular intervals. Features such as
crack stoppers and load transfer between components should also be considered during
design, so that the structure can sustain loading for a period of time before the damage is
detected. Today, all Boeing jets are designed with the fail-safe principle, and it is also used
outside of the aerospace industry[2, 13].

Increasing Fatigue life

Increasing the lifetime of a component or structure that is subject to fatigue can be


achieved by both proper design and manufacture, and by improvement of the conditions it
is subjected during service.

It has been observed that increasing the strength of a metal or alloy improves its
fatigue behaviour, since the cracks propagate by an elastic micromechanism[3]. As the
cracks initiate on the surface, it is usually enough to only increase the strength of the
material only near the surface. A number of hardening processes that are employed in
order to increase a metal’s fatigue behaviour are age hardening[14], cold working[15],
nitriding [16] and carburizing[17].
P a g e | 11

Another common strategy is the introduction of compressive stresses on the surface


of a metal, which counteract the tensile stresses that lead to crack initiation. This is achieved
with shot peening: impacting the surface with tiny balls or particles with force, leading to
plastic deformation and residual stresses[18]. Roller burnishing, which is pressing the
component with highly polished steel rollers in order to plastically deform its surface, can
also be employed[19].

Improving the surface smoothness and as a result reducing grooves where stress
concentration can lead to crack initiation is also a way to increase fatigue life. This can be
done with mechanical polishing[19] or chemically with elecropolishing[20].

Finally, as microstructure affects crack propagation (as discussed earlier), altering


material properties such as grain size, purity, and texture can lead to a significant increase in
fatigue life.

Monitoring component fatigue during service

Evaluating the condition of a component during service is another important part of


avoiding failure due to fatigue. The techniques that are employed do not destroy the
component, and are known as non-destructive testing (NDT). They include the use of
ultrasonic [21], infrared light (as the crack has an increased heat absorption)[22], dye
penetrants[23], magnetic particle inspection[23], radiography[23], and eddy currents[24].
P a g e | 12

Diagram 12: Detection of surface cracks using dye penetrant

References

1. STP, A., E1150 Annual Book of Standards. ASTM, Philadelphia, PA, 1996. 3: p. 740-49.
2. Stephens, R.I., et al., Metal fatigue in engineering. 2000: John Wiley & Sons.
3. Handbook, A., Vol. 19 “Fatigue and Fracture”. ASM International, 1996.
4. Polák, J. and J. Man, Fatigue crack initiation–the role of point defects. International Journal
of Fatigue, 2014. 65: p. 18-27.
5. Gueler, S., M. Schymura, and A. Fischer, Austenitic high interstitial steels vs. CoCrMo–
Comparison of fatigue behavior. International Journal of Fatigue, 2015. 75: p. 145-152.
6. Silvestre, E., et al., Comparison of the hardening behaviour of different steel families: From
mild and stainless steel to advanced high strength steels. International Journal of Mechanical
Sciences, 2015. 101: p. 10-20.
7. Milella, P.P., Fatigue and corrosion in metals. 2012: Springer Science & Business Media.
8. Callister, W.D. and D.G. Rethwisch, Materials science and engineering: an introduction. Vol.
7. 2007: Wiley New York.
9. Chan, K.S., Roles of microstructure in fatigue crack initiation. International Journal of Fatigue,
2010. 32(9): p. 1428-1447.
10. Ishihara, S., Z. Nan, and T. Goshima, Effect of microstructure on fatigue behavior of AZ31
magnesium alloy. Materials Science and Engineering: A, 2007. 468: p. 214-222.
11. Bathias, C., There is no infinite fatigue life in metallic materials. Fatigue and Fracture of
Engineering Materials and structures, 1999. 22(7): p. 559-566.
12. Sun, Q., H.-N. Dui, and X.-L. Fan, A statistically consistent fatigue damage model based on
Miner’s rule. International Journal of Fatigue, 2014. 69: p. 16-21.
13. Goranson, U.G., Fatigue issues in aircraft maintenance and repairs. International Journal of
Fatigue, 1997. 19(93): p. 3-21.
14. Sadeler, R., et al., Improvements of fatigue behaviour in 2014 Al alloy by solution heat
treating and age-hardening. Materials & design, 2004. 25(5): p. 439-445.
15. De Matos, P., et al., Analysis of the effect of cold-working of rivet holes on the fatigue life of
an aluminum alloy. International Journal of Fatigue, 2007. 29(3): p. 575-586.
P a g e | 13

16. Genel, K., M. Demirkol, and M. Çapa, Effect of ion nitriding on fatigue behaviour of AISI 4140
steel. Materials Science and Engineering: A, 2000. 279(1): p. 207-216.
17. Tokaji, K., K. Kohyama, and M. Akita, Fatigue behaviour and fracture mechanism of a 316
stainless steel hardened by carburizing. International journal of fatigue, 2004. 26(5): p. 543-
551.
18. Llaneza, V. and F. Belzunce, Optimal Shot Peening Treatments to Maximize the Fatigue Life
of Quenched and Tempered Steels. Journal of Materials Engineering and Performance: p. 1-
10.
19. Hilpert, M. and L. Wagner, Corrosion fatigue behavior of the high-strength magnesium alloy
AZ 80. Journal of Materials Engineering and Performance, 2000. 9(4): p. 402-407.
20. Lopes, H.P., et al., Effects of electropolishing surface treatment on the cyclic fatigue
resistance of BioRace nickel-titanium rotary instruments. Journal of endodontics, 2010.
36(10): p. 1653-1657.
21. Jhang, K.-Y., Nonlinear ultrasonic techniques for nondestructive assessment of micro damage
in material: a review. International journal of precision engineering and manufacturing,
2009. 10(1): p. 123-135.
22. Kubiak, E.J., Infrared detection of fatigue cracks and other near-surface defects. Applied
Optics, 1968. 7(9): p. 1743-1747.
23. Willcox, M. and G. Downes, A brief description of NDT techniques. Toronto: NDT Equipment
Limited, 2003.
24. Zilberstein, V., et al., MWM eddy current sensors for monitoring of crack initiation and
growth during fatigue tests and in service. International journal of fatigue, 2001. 23: p. 477-
485.

You might also like