You are on page 1of 14

Johannes Homan, MSc

Fatigue life prediction is a common task for engineers. How do we make


sure that we are doing the right thing? There are many codes and FE tools
available for fatigue analysis. Are those tools doing it right?
This ebook presents an overview of common mistakes made in fatigue
analysis.

1. Misunderstanding fatigue
Crack initiation and crack growth are driven by cyclic loads, constant
stresses will not cause fatigue crack initiation.

2. Confusing stress amplitude and stress range


In general, S-N curves are given with stress amplitudes vs. number of
cycles to failure. The stress amplitude is defined as half the stress range.
For welded joints, S-N curves are usually given with the stress range. An
error by factor 2 in stresses is therefore easily made.
Sometimes, the maximum stress level instead of the stress amplitude is
used in S-N curves.

3. Ignoring mean stress and residual stresses


Although fatigue life is determined by the cyclic character of stresses (i.e.
stress amplitude), the mean stress has an effect as well. An increasing
mean stress is unfavourable for the fatigue life, a decreasing mean stress
is favourable.
Note that residual stresses at the critical location (pre-stresses, assembly
stresses, etc.) have the same effect as mean stresses of the cycle.

March 2018 2 Doc. 2018-002/jh


4. Deriving fatigue data from static properties
The fatigue strength depends on much more factors than the static
strength. Surface conditions, stress concentrations, component size,
loading type, mean stress level, environment, etc., have a large impact on
the fatigue properties. For example, a high strength steel with a very sharp
notch in a corrosive environment under variable amplitude loading will
have a fatigue limit of approximately 1% of the tensile strength whereas
a mild steel, soft condition, un-notched, inert atmosphere, constant
amplitude loading may have a fatigue limit of close to 70% of the tensile
strength.
Nevertheless, it is often suggested that the fatigue limit for the case Kt=1
and R=-1 can be derived from the tensile strength of the material. The
figure below shows that there is indeed a trend in the relation fatigue
strength – tensile strength, but also that the amount of scatter is very
large. Such a relation should therefore only used as an indication, but
should certainly not be used if some degree of accuracy is required.

March 2018 3 Doc. 2018-002/jh


5. Applying mean stress correction on stresses
instead of fatigue limit
S-N curves are often given for a mean stress level of zero (R=-1). Mean
stress corrections (Gerber, Goodman, Schütz, FKM) are available to
estimate the fatigue data (actually the fatigue limit) for mean stress levels
other than zero. Note that the mean stress not only affects the fatigue
limit, but also the slope of the S-N curve. That implies that applying a
mean stress correction (MSC) on the applied stresses instead of the fatigue
data will give a different (i.e. incorrect) result in the analysis. In case MSC
applied on stresses does give the same result, some other mistake has
been made, e.g. not adjusting the S-N curve slope.

6. Assuming welds behave like base material


Welding ruins the carefully built microstructure of the base material but
only locally at the weld. In the base metal the properties are determined
by the processing of the material. In the weld zone the material processing
is overruled by the welding process, resulting in a (fast) solidified structure
with decreased properties. The figure below shows the large reduction in
fatigue limit compared to base material.

March 2018 4 Doc. 2018-002/jh


Another reason for reduced fatigue strength in welds are welding defects
(porosity/voids, slag inclusions, incomplete fusion & penetration and
residual stresses).
The detrimental effect of welding is much larger than the effect of stress
concentrations. This is illustrated in the graph below, showing S-N curves
for an unnotched specimen, a notched specimen and a specimen with a
welded gusset.

7. Incorrect cycle counting


Just counting the consecutive cycles like in the figure below may lead to a
large overestimation of the fatigue life.

March 2018 5 Doc. 2018-002/jh


The best way to count cycles is “Rainflow Counting”. This method roughly
implies that small cycles are taken from large cycles and counted
separately, whereas the remaining large cycles are counted as well.

8. Underestimating small cycles


Small cycles (i.e. cycles with a small amplitude) lead to longer fatigue lives
than large cycles (i.e. cycles with a large amplitude). This fact may lead to
the incorrect conclusion that small cycles can be neglected. However, in a
spectrum the number of small cycles is often much larger than the number
of large cycles. If so, the small cycles do give a significant contribution to
the damage accumulation.
It is sometimes even thought that small cycles can be neglected, resulting
in a small number of large cycles, which situation then erroneously is
interpreted as LCF. Note that LCF corresponds with cyclic plastic
deformations, not with small number of occurrences of large (elastic)
cycles.

March 2018 6 Doc. 2018-002/jh


9. Assuming simplified loads
When load spectra are compiled, often only functional loads are taken into
account. Disturbances (usually of stochastic nature) however are often
overlooked or underestimated. In the figure below an example is given of
a part of a spectrum with functional loads and disturbances superimposed
on the functional load. The disturbances lead to an amplitude increase of
the largest cycle (rainflow counting) and to an increased number of small
cycles.

10. Trusting the Miner rule


Damage accumulation according to the Miner rule is using this equation:
ni
N =1
i

With ni being the number of occurrences of cycle i and Ni being the fatigue
life that corresponds with the stress amplitude of cycle i. All cycles above
the fatigue limit have a finite life N and therefore give a contribution to
the damage accumulation. However, as soon as those cycles have created
some damage (micro crack) also cycles below the fatigue limit will start

March 2018 7 Doc. 2018-002/jh


contributing to damage accumulation. This contribution is not included in
the Miner rule. To account for the damage accumulation by small cycles,
some options are available. One option is to ignore the fatigue limit and
extrapolate the S-N curve below the fatigue limit, i.e. applying Sak·N=c
also for Sa<Sf. This option gives very conservative results. Another option
is the Haibach extrapolation, i.e. using Sa2k-1·N=c for Sa<Sf. This option is
rather arbitrary and not accurate.

A better solution is the relative Miner rule:


n
N =D
The failure criterion D is then smaller than 1 an depends on the actual
load sequence. Unfortunately there is hardly any data available on D in
literature.
Another issue that make the Miner rule notoriously inaccurate is the effect
of notch root plasticity. High stress levels cause residual stresses at the
notch which have an effect on the actual mean stress at the notch in the
subsequent cycles.

March 2018 8 Doc. 2018-002/jh


11. Using fixed S-N curve slope for local
stresses
The common approach for fatigue life analysis is using S-N curves based
on nominal stresses (i.e. average stresses in the net section). Depending
on the severity of the stress concentration (Kt) in the structure, a S-N curve
is then selected that corresponds with the actual structure. In the diagram
below a few S-N curves are shown for different Kt values. The larger the
Kt, the lower the fatigue limit and the steeper the S-N curve (smaller value
of k in the S-N curve equation (Sak·N=c).

In case the fatigue analysis is performed using local stresses, the S-N
curves based on nominal stresses cannot be applied, instead S-N curves
based on local stresses must be used. In the diagram below, such a set of
S-N curves is shown. These curves are derived from the S-N curves based
on nominal stresses shown above, by multiplying the nominal stresses by
Kt to obtain the local stresses. The S-N curves based on local stresses
show different fatigue limits and different gradients for different Kt values.

March 2018 9 Doc. 2018-002/jh


Always be aware that a local stress can be any combination of global stress
and Kt. Additionally, the gradient depends on the stress ratio as well.

12. Using S-N data for LCF


In the LCF (Low Cycle Fatigue) region, fatigue is driven by cyclic strains.
The material behaviour in that region is dominated by plastic deformation,
non-linear relation between strains and stresses. An S-N curve describes
fatigue in the HCF (High Cycle Fatigue) region, i.e. elastic material
behaviour and a linear relation between strains and stresses.

March 2018 10 Doc. 2018-002/jh


13. Using -N data for HCF
For HCF, sometimes the elastic part of the -N curve is used. It is argued
that because of the linear elastic material behaviour in that region, the -N
curve is in fact a S-N curve. Since the strains are local strains, and the -N
curve is based on unnotched specimens, this approach leads to the same
problems as described in section 11.
Furthermore, the position of the elastic part of the -N curve is partly
determined by the fatigue strength coefficient ’f. This coefficient is
derived from a static tensile test and has nothing to do with high cycle
fatigue.

14. Using equivalent stresses like Von Mises


Fatigue damage is propagating perpendicular to the largest principal stress
range, therefore this stress range determines fatigue behaviour. Principal
stresses in tension in the other directions have hardly any influence on the
crack growth, these stresses do not affect the shear stress in the activated

March 2018 11 Doc. 2018-002/jh


slip planes. In case of shear (2D stress state with bi-axiality ratio of, or
close to, -1), fatigue data for shear should be used.
3D stress states are hardly ever relevant for fatigue, crack always start at
a free surface. Even with sub-surface initiation it can be argued that the
stress state is 2D, cracks start at inclusions or voids.
If there is a 2D or 3D stress state with varying largest principal stress
direction, it is thought sometimes that using equivalent stresses is
attractive. This is not the case, equivalent stresses have no direction and
certainly not a varying one. The best approach would be the “Critical Plane
Approach”, i.e. analysing different crack growth directions (planes) and
using for each plane the stress components perpendicular to that plane.
Note that such an approach must be performed twice; viz. also for shear
stresses.

15. Using scatter data from laboratory tests


for actual service applications
Fatigue always shows a lot of scatter. This scatter is caused by (1)
variability in material properties, (2) variability in surface conditions
(process/manufacturing), (3) variability in loading, and (4) variability in
environment. S-N data obtained from laboratory tests only cover the first
two causes. If the S-N data just exists of the fatigue limit being a function
of the tensile strength, only variability of material properties is taken into
account.
The standard deviation (log-normal distribution of N) for laboratory S-N
data is lower than 0.25 (often between 0.05 and 0.15). For scatter in
service, typical values for the standard deviation range from 0.25 to 0.35.

March 2018 12 Doc. 2018-002/jh


So if an analysis based on laboratory scatter data gives a fatigue life with
a certain probability of failure, this result is inaccurate.

16. Ignoring surface conditions


Surface conditions have a large impact on fatigue behaviour but are easily
neglected. S-N curves are often established using nicely manufactured
specimens. Actual production of your components may be completely
different. This difference should be taken into account.

March 2018 13 Doc. 2018-002/jh


Do you want to know more about fatigue analysis?

In this book, guidelines are given for fatigue life estimations


and analysis. A known quote from Arthur Schopenhauer is:
“Wo das Rechnen anfängt, hört das Wissen auf”. One may
translate this to: “Where calculating starts, understanding
ends”. The objective of this book is to give the reader
background information and knowledge on fatigue analysis
methods (i.e. understanding the calculation). It is therefore
neither a text book nor a recipe book, but the missing link in-
between.

https://www.fatec-engineering.com/publications/

About Johannes Homan


Johannes has a background in aerospace engineering and has specialized himself in
fatigue and damage tolerance. After his graduation at Delft University of Technology
(faculty of Aerospace Engineering) Johannes started working at Fokker Aircraft as
Fatigue & Damage Tolerance specialist.
In 1996 Johannes started his own firm: Fatec Engineering and works in this firm since
then. During that period, Johannes was also active as researcher and assistant
professor at Delft University of Technology (faculty of Aerospace Engineering) and as
researcher at GTM Advanced Structures.
LinkedIn: https://www.linkedin.com/in/johanneshoman/

Rubenslaan 127
2661 RV Bergschenhoek
The Netherlands
Phone: +31 10 5299713
Email: johannes.homan@fatec-engineering.com
https://www.fatec-engineering.com

March 2018 14 Doc. 2018-002/jh

You might also like