You are on page 1of 28

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/274613109

Fatigue behavior of the aeronautical AlLi (2198)


aluminum alloy under constant amplitude
loading
Article in International Journal of Fatigue November 2013
DOI: 10.1016/j.ijfatigue.2013.07.009

CITATIONS

READS

18

192

4 authors, including:
Nikolaos D. Alexopoulos
University of the Aegean
59 PUBLICATIONS 562 CITATIONS
SEE PROFILE

All content following this page was uploaded by Nikolaos D. Alexopoulos on 11 May 2015.
The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.

International Journal of Fatigue 56 (2013) 95105

Fatigue behaviour of the aeronautical Al-Li (2198) aluminum alloy under


constant amplitude loading
Nikolaos D. Alexopoulos1*, Evangelos Migklis1, Antonis Stylianos1 and Dimitrios P. Myriounis2
1

Department of Financial Engineering, University of the Aegean, 41 Kountouriotou str, 82100, Chios, Greece

Materials and Engineering Research Institute, Sheffield Hallam University, Howard Street, S11WB Sheffield, UK

Abstract
Tensile and fatigue mechanical behaviour of wrought aluminum alloy 2198-T351 is examined and
compared against 2024-T3 that is currently used in aero structures. Experimental fatigue tests were
carried out under constant amplitude stress ratio R = 0.1 and respective stress - life (S-N) diagrams were
constructed for both alloys. Fatigue behaviour of both alloys is described with varying parameters being
the percentage of fatigue life as well as the effect of maximum applied stress as a function of ultimate
tensile strength. It was found that fatigue endurance limit of AA2024-T3 is approximately 40 % below its
yield stress, while only 9 % below for the AA2198-T351. The latter was found to be superior in the high
cycle fatigue and fatigue endurance limit regimes, especially when considering specific mechanical
properties. Absorbed energies per fatigue cycle as well as dynamic stiffness of the fatigue hysteresis loop
were calculated and plotted against the number of fatigue cycles and with varying maximum applied
stress; both parameters are continuously decreasing due to the combination of hardening effect and microcracking in AA2024-T3, while this was the case only for the high applied stresses regime in AA2198T351. Cyclic stress strain (CSS) curves were constructed and proved that work hardening exponent of
AA2198-T351 is substantially decreasing with increasing fatigue life.

Keywords: aluminum alloys; fatigue; fracture; tension; specific properties; fractography

1. Introduction
The aerospace industry is always demanding innovative, lighter aluminum alloys with improved
mechanical properties. Aluminum lithium (Al-Li) alloys offer great advantages for use in aerostructures
through density reduction, stiffness increase, increases in fracture toughness and fatigue crack growth
resistance as well as enhanced corrosion resistance. The addition of Li, results in the reduction of the

*1Corresponding author: e-mail: nalexop@aegean.gr

International Journal of Fatigue 56 (2013) 95105


weight of the alloy, since for each 1 wt% Li added to Al, density is reduced by 3 % and modulus of
elasticity is increased by almost 6 % [1]-[3]. The previous generations of Al-Li alloys (e.g. 2090, 8090
and 2091) exhibited significant in-plane and through-thickness anisotropy in mechanical properties.
Nevertheless, they were responsible to unpredictable failures during manufacturing (e.g. during cold hole
expansion) and made the design of aerostructures very complicated since they exhibited low shorttransverse fracture toughness [2].
The third generation of damage tolerant Al-Li alloys has been developed in the last decade by
taking into account all above disadvantages; key alloy design principles led to the successful development
and commercialization of the third generation Al-Li alloys with highly desirable combinations of
mechanical properties [4]. AA2198 is such an Al-Li alloy that is known to present extremely high
strength levels due to the precipitation hardening mechanisms. It has a wt% Cu composition ranging from
2.9 to 3.3 % and respective of Li from 0.9 to 1.1 %. The third-generation Al-Li alloys have relatively
lower Li level when compared with first- and second-generation alloys. Their final precipitation state is
reached via a complex sequence that can involve notably solute clusters, GP zones, , T1 and phases
[4]-[6]. Addition of Cu and Mg play an important role by promoting the formation of T1 (Al2CuLi)
metastable strengthening phase [7], at the expense of (Al3Li), which is generally absent for Li levels <
1.3 wt% [11]. Other phases have been reported to form in these alloys, including / (Al2Cu) [8], [9] as
well as and S [6], [8], depending on the alloys composition. From all the above age-hardening phases,
the T1 hexagonal platelets seems to have the highest metallurgical interest, since they are lying on the {1
1 1} matrix planes and are very effective at blocking slip [10],[11].
At the same time, the introduction of the welding process into manufacturing of aerostructures
favors the further reduction of weight and cost savings by replacing riveting and fastening procedures.
Hence, the innovative alloys developed for the aerospace industry should be capable to be efficiently
welded. Despite the problems that may arise with fusion welding such as porosity arising from the
addition of lithium [12] and solidification cracking [4],[13], Al-Li alloys are considered as highlyefficient weldable alloys, e.g. [14]-[16]. Since only a 10 to 15 % decrease in fatigue life of the welded
specimens is noticed [17], Al-Li alloys have already found application in the aerospace sector [2].
The quasi-static mechanical properties of AA2198-T351 are scarcely reported in the open
literature. For example, Chen et al. [18] performed tests on two different heat treated AA2198 (namely
T351 and T851) and investigated their plastic and fracture behavior. Seglich et al. [19],[20] investigated
experimentally and analytically the anisotropic deformation of 2198-8 occurred during mechanical
loading with and without the presence of artificial notches. Fatigue behaviour of aluminum alloys is very
critical to such structures, since more than 40 % of damages repaired during aircraft maintenance
originate from metal fatigue [21]. Fatigue behavior of uncracked and cracked specimens of AA2024-T3
has been examined in depth and reported extensively in the open literature during the last decades, e.g.
2

International Journal of Fatigue 56 (2013) 95105


[22]-[27]. Other researchers focused on the characterization of AA2024 fatigue behavior and the effects
of age-hardening conditions [28]-[30].
Welding efficiency of AA2198 seems to be one of its advantages when compared to the
conventional AA2024 that is currently used in aerostructures. In order the Al-Li alloy to replace the
conventional one in aircraft structures, it has to be proven that its mechanical behaviour and damage
tolerance capabilities are at least equal or superior to its predecessor. Nevertheless, the fatigue mechanical
performance of the AA2198 is scarcely reported. The only data that the authors could find in the open
literature was the reference S-N curves of AA2198-T8 [14] and T851 [17] conditions, respectively. Both
S-N fatigue curves of reference coupons were plotted against the respective curves of friction stir welded
specimens, in order to understand and to demonstrate the mechanical behaviour of the joints with the
innovative welding process.
As AA2198 is supposed to replace AA2024 in aerostructures designed with the damage tolerance
philosophy, the authors in the present work will try to report and compare their fatigue mechanical
behaviour under constant amplitude loadings. To this end, a comparison of their tensile mechanical
behaviour is needed first; S-N fatigue curves will be constructed and individual characteristics such as
energy absorbed per fatigue hysteresis loop, residual stiffness (modulus of elasticity) and cyclic stress
strain curves will be calculated per different alloy as a function of maximum applied stress and
percentage of fatigue life, respectively. The results will highlight whether AA2198-T351 has overall
higher fatigue mechanical behaviour capabilities than the conventional AA2024-T3.

2. Materials and specimens


The materials used for the present investigation were AA2024-T3 and AA2198-T351 wrought aluminum
alloys that were received in sheet form with nominal thicknesses of 3.2 mm and 1.6 mm, respectively.
The sheets were provided by Hellenic Aerospace Industry with geometrical dimensions of 35 x 50 cm,
while their thickness is typical for such aerostructures. Typical chemical composition of the delivered
sheets can be seen in Table 1. Tensile and fatigue specimens were machined from the longitudinal (L)
orientation only and according to the specifications ASTM E8M and E466, respectively. The geometrical
dimensions of the machined specimens can be seen in Figure 1. Ten (10) tensile specimens were
machined from each aluminum sheet, while more than twenty (20) fatigue specimens were machined
from the same sheets.

3. Experimental procedure
3

International Journal of Fatigue 56 (2013) 95105


Tensile mechanical tests were carried out in a servo-hydraulic Instron 100 kN testing machine. An
external extensometer was attached to the specimen surface and at the reduced cross-section gauge length.
Tensile tests were carried out according to ASTM E8 specification and the strain rate was kept constant
and equal to 0.833 sec-1. A data logger was used during all the experiments and the values of load,
displacement and axial strain were recorded in a computer.
Fatigue tests were carried out in the same testing machine, Figure 1; all tests were performed with
load command and by using sine wave load. Load (or stress) ratio was kept constant for fatigue tests and
equal to R = 0.1, which is the typical waveform for testing aeronautical materials, since it better simulates
the true loadings in an aerostructure. An Instron dynamic extensometer was attached to the specimens
surface. Fatigue load frequency was 10 Hz for all tests; hysteresis loop of the specimen were recorded as
well as the values of load, displacement and axial strain were recorded and stored in a computer.
Fractured surfaces of the fatigued specimens were examined with the aid of a FEI NOVA 200
NanoSEM scanning electron microscope (SEM).

4. Mechanical test results


4.1 Tensile test results
Experimental tensile test results of the investigated alloys can be seen in Table 2 in the form of average
values and standard deviation, while typical engineering stress - strain tensile curves are presented in
Figure 2. Al-Li alloys modulus of elasticity was evaluated around 75 GPa, that is more than 7 % higher
than AA2024. The conventional alloy exhibits essential higher strength values than the respective of
AA2198-T351. For example, the latter exhibits lower yield stress and tensile strength by 33 % and 23 %,
respectively, Figure 3. AA2198-T351 seems to present another disadvantage of lower elongation at
fracture Af values; it is calculated to be almost 21 % lower than the conventional alloy. The same applies
to the other ductility measure; lower strain energy density W values were recorded than 2024-T3, Table 2.
The lower strength values of AA2198-T351 might be a real disadvantage when used in real
aerostructures. However, artificial aging of the precipitated AA2198 might be the solution to this
problem. Figure 2 shows a typical tensile curve of the artificially aged Al-Li alloy at the T851 condition.
It can be seen that its strength disadvantage has been clearly diminished, as the aged alloy presents now
higher yield strength values Rp = 432 MPa which is even higher that the respective of AA2024-T3.
Tensile strength values are almost the same for both alloys, while ductility decreased, taking values
around Af = 12 % that corresponds to almost 60 % of the conventional alloy, Figure 3. Nevertheless, it is
clear that due to artificial aging, alloy can withstand the increase in strength on the expense of ductility
4

International Journal of Fatigue 56 (2013) 95105


and vice versa, which seems to be a real advantage when trying to tune the mechanical properties of the
structure to the required.
To characterize the plastic deformation of the material, work hardening exponent n is usually
employed and it represents more-or-less the rate of plastic hardening. Work hardening exponent n was
calculated for all performed tensile tests. Hollomon equation was suggested to be valid from the
beginning of plastic deformation and up to the formation of necking, as pointed out by ASTM E646:

= n ,

(1)

where is the true stress, is the true strain, K is the strength coefficient and n is the work hardening
exponent (stage II linear regime). AA2024-T3 presented values n = 0.1064, while AA2198-T351
presented high hardening exponent values, taking as an average that n = 0.1486. As expected, with
increasing artificial aging treatment (T851), hardening exponent significantly decreases and took values
of approximate n = 0.0458. Hence, aging at T851 condition brings the alloy at very low hardening
exponent that is even lower than the reference 2024-T3 alloy.

4.2 Fatigue stress life curves


Fatigue stress - life (S-N) curves of the investigated alloys were constructed from the available performed
fatigue constant amplitude tests. Figure 4 shows the respective semi-logarithmic S-N curve of AA2024T3 and the experimental data can be seen as semi-filled circles. In total twenty one fatigue tests had been
performed; experimental fatigue data were plotted in the stress - lifetime diagram. Three different regimes
can be noticed: (i) high applied stresses (> 450 MPa or less than 10 kilocycles) that corresponds more-orless to the low cycle fatigue regime (LCF), (ii) medium applied stresses which is evident of the High
Cycle Fatigue (HCF) regime that lies in the range between 250 and 450 MPa and finally (iii) low applied
stresses or fatigue endurance limit, which is below the threshold of 240-250 MPa maximum applied stress
for stress ratio of R = 0.1. Fatigue testing of specimens was interrupted at the fatigue endurance limit
regime and after completing 5 x 106 fatigue cycles.
S-N fatigue data from other articles of the open literature [23],[26],[27] were also extracted and
plotted in the same diagram. Fatigue S-N data were extracted with different stress ratio, namely R = 1
(Whler curve) or even R = 0.33. To this end, the available literature data were converted to the same
stress ratio of R = 0.1 in order to be compared with the experimental results. This was achieved by
employing the Gerbers relation, as [31]:
2
2
o

+
= 1 or = cr 1 o ,

cr Rm
Rm

(2)

International Journal of Fatigue 56 (2013) 95105


where is the alternating-stress amplitude; cr is the completely reversed stress amplitude for a given life
(i.e., 106, 105, etc.); Rm is the ultimate tensile strength of the material and o is the mean stress. The use of
the Gerbers formula was preferred than Soderbergs and Goodmans rules, since the first one gives
conservative results and the second despite that it works in brittle materials, it is conservative to ductile
materials [32], which is the case in the present study.
As it can be seen in Figure 4, all AA2024 S-N curves seems to have the same generic fatigue
behaviour with the exception of the fatigue endurance limit regime. Nevertheless, given the typical
experimental scatter of fatigue testing as well as the different thicknesses (3.2 mm against thicknesses
ranging from 5 to 12 mm) and temper (T3 against T351) used, these differences between the experimental
data and the results from bibliography are considered acceptable.
The fatigue test results in terms of S-N curve of AA2198-T351 can be seen in Figure 5. In total
eighteen fatigue tests were carried out on specimens that were cut in the longitudinal (L) orientation. As
can be seen in the Figure, the plotted curve presents a linear regime from 320 until 260 MPa, which is
typical of the high cycle fatigue regime. From 260 MPa and below, the S-N curve tends to stabilize,
which is typical for the fatigue endurance limit regime and the conventional fatigue endurance limit is
approximately evaluated to be 240 5 MPa. For higher applied stresses than 320 MPa the regime
corresponds to LCF with the exception that the test command was load-controlled than strain-controlled,
which is the case in a proper low cycle fatigue test. In the diagram, three different specimens were
selected to be examined in scanning electron microscope for fracture analysis; one (No 11) at the
endurance limit and two other (No 4 and 5) at the high cycle fatigue regime that will be explained in the
fractography section.
Literature fatigue data for AA2198-T351 seems to be very rare in the open literature, as the only
record [14] the authors could find was for the same alloy with different temper condition (T851 instead of
T351), different thickness (5 mm instead of 1.55 mm) and different stress ratio (R = 0.33 instead of R =
0.1). The latter was corrected according to the Gerbers rule and by employing eqn.(2) twice; the first was
to convert the data to from R = 0.33 to 1, while the second was to convert them from R = 1 to 0.1
which is the stress ratio used in the present study. Literature and experimental data do not argue precisely
in terms of fatigue endurance limit due to above described reasons; however it is evident that both curves
follow the same trend.
To exploit the available experimental fatigue mechanical behaviour, S-N curve of AA2198-T351
was compared against the respective of AA2024-T3. A direct comparison of the two S-N curves in the
previous figures would definitely favor the conventional alloy, since it has the capability to undertake
higher applied load (stress) spectra. They both seem to present the same fatigue endurance limit, around
240 MPa. Surprisingly, the endurance limit is almost 40 % below AA2024 yield stress Rp, while for 2198T351 is in-between 7 to 10 % lower than the respective Rp value. This would enhance lowering the safety
6

International Journal of Fatigue 56 (2013) 95105


factors used in aerostructures that currently is a percentage of the alloys yield stress. However, this
finding needs extensive research to support why the endurance limit is so close to the yield stress value
and to distinguish the role of Li and its precipitates to the nucleation and propagation of micro-cracks.
One of the advantages of the newly-developed alloy is its lower density; as the design office in
aircraft structures uses the specific mechanical properties of materials, the available fatigue experiments
were plotted in the diagram of Figure 6 with the vertical axis being the specific maximum applied stress
(i.e. max divided by density). For the high applied stresses regime, AA2024 clearly presents better fatigue
mechanical behaviour as higher specific stresses can be applied to specimens. This is also the case to the
first half HCF regime that is marked in the diagram, where for the same applied specific stress AA2024
presents higher number of fatigue cycles to fracture. However, after the marked threshold of approximate
75 kilocycles (lower specific stresses), this trend is reversed and AA2198 presents more fatigue cycles to
fracture for the same specific stress. In detail, the new alloy presented 25 % to 60 % higher fatigue cycles
to fracture for applied specific stresses of 110 and 100 (MPa.m3)/kg, respectively. To the authors
knowledge, this finding is reported for the first time in the open literature. Again, the very good fatigue
behaviour of AA2198-T351 can also be seen for the regime of fatigue endurance limit, where it presents
substantially higher (3%) fatigue endurance limit than AA2024-T3.

5. Analysis
5.1 Residual Stiffness
A fatigue hysteresis loop was formed during the loading and unloading branches per every fatigue cycle.
It is well-known that the formation of these hysteresis loops is a function among all of the material
(chemical composition and temper), of the applied stresses and of the materials fatigue life. For the
purpose of the present work, the moduli of elasticity of all specimens were calculated per every fatigue
unloading branch and for all stages of fatigue life till fracture. The unloading branch was simulated by a
linear curve equation, its slope giving the unloading stiffness, Figure 7. In the current work the modulus
of elasticity per different stages of fatigue life will be denoted as residual stiffness Eun. This term was
better selected since during the lifetime of fatigue test several irreversible phenomena like crack
nucleation and growth are taking place. Hence, since the term of modulus of elasticity is linked with the
crack-free and damage-free material, the term of stiffness is preferred since during the experiment of
fatigue, micro- and macroscopic kinds of damage had been developed in the specimens. Finally, the
branch of unloading was selected by the authors to calculate its stiffness since it contains no parasitic
effects that might take place during the loading branch, e.g. inertia of the loading frame, etc. The residual
7

International Journal of Fatigue 56 (2013) 95105


stiffness was calculated per every single fatigue experiment and for different stages of its fatigue life as
the slope of the nominal stress strain unloading branch and was calculated in GPa.
The calculated values of stiffness of the two alloys were plotted in the diagrams of Figure 8, where
the horizontal axis represents the percentage of fatigue life (or damage) in a linear scale, while the vertical
axis represents the values of residual stiffness. To demonstrate the percentage of fatigue life, the linear
rule of Miner was chosen, mainly for simplicity reasons. It well known that the Miners rule (Palagren
Miners rule) [32] does not work well in irregular fatigue spectra and does not take into account the
fatigue loading history, however it works very good with single-block constant amplitude fatigue tests
which is the case of the present work. Therefore, the term of percentage of fatigue lifetime was calculated
as the % ratio of the current fatigue cycle ni to the fracture cycles Nf of the same specimen:
m

ni

N
i =1

= D.

(3)

Figure 8(a) shows the calculated results of AA2024-T3 dynamic stiffness for the different percentages of
its fatigue life and for different applied maximum stresses. In the diagram, the different curves correspond
to different maximum applied fatigue stress, while marked is its respective percentage when divided to
ultimate tensile strength (normalized stress). Dynamic residual stiffness takes values ranging from 78 GPa
for the smallest applied stress of 220 MPa till approximately 50 GPa for the highest applied stress. For all
performed fatigue tests below yield stress (Rp = 391 MPa), dynamic stiffness takes values around the
respective quasi-static value. For the very first stages of fatigue life and up to approximate 15 % fatigue
time, it can be seen that the stiffness is increasing highly dependant on maximum stress. This
phenomenon can be noticed clearly for the high applied stresses, e.g. for max = 490 MPa and to a lesser
extend to others due to resolution. Of course, this is evidence of hardening effect that will be discussed in
the following. In addition, a minor reduction of the residual stiffness in the high percentages regime of
fatigue damage can be clearly observed at the experimental curves, e.g. for max = 280, 350, 440 and 480
MPa and despite the low resolution of the figure. Such a decrease in stiffness is definitely evidence of
crack propagation stage (stage III), as the nucleated fatigue crack propagates. More details will be given
in the very next paragraph.
Figure 8(b) shows the calculated AA2198-T351 results; the same trend also applies for this case.
Stiffness is increasing with decreasing maximum fatigue stress and the results differ when compared to
with AA2024 only to the smaller span (ranging from 82 GPa till 67 GPa). However, these results are in
accordance with the AA2198-T351 higher modulus of elasticity evaluated by the quasi-static flow curve.
In addition, the curves seem to present a smooth increase at the early stages of fatigue life that implies
hardening. A small decrease is noticed at the last stages of fatigue life that can be clearly seen at stresses
higher than yielding, e.g. at max = 275, 285 and 315 MPa. This was also noticed for the AA2024-T3 case
and it is evidence of the specimens transition from stage II to stage III, i.e. unstable crack propagation.
8

International Journal of Fatigue 56 (2013) 95105


To this end, all the available fatigue stiffness data of both alloys were analyzed separately. For
every fatigue experiment, the transition point (percentage of fatigue life) where the stiffness first began to
decrease (Decrease Point Dp) was evaluated as a function of maximum applied stress. The results can be
seen in Figure 9 in terms of normalized maximum stress max to ultimate tensile strength Rm (x-axis) and
percentage of fatigue life where the decrease of stiffness initiated Dp (y-axis). Yield stresses of both alloys
are also marked in the diagram; for the case of AA2024-T3, stage III seems to initiate at almost 80 % of
fatigue life at stresses below yield strength marked in blue circles (~ 57 % of normalized max/Rm) and is
decreasing with increasing applied stress. For the case of max > Rp, then there exists a monotonic
(probably non-linear) decrease of this initiation point, meaning that stage III occurs from almost 50 % of
fatigue life at max = 96 % of Rm. One AA2198-T351 fatigue test was performed below yield stress, e.g. at
65.7 % of Rm; many experiments have been performed for maximum stresses higher than yield stress and
a linear approximation could be established with increasing applied stress. The difference between the
two alloys is evident; for the case of fatigue stressing at approximately 90 % of Rm, stage III occurs in 20
and 60 % of fatigue life for the 2198 and 2024, respectively. But the impression that AA2024-T3 seems
to have a better fatigue behaviour in high applied stresses regime was evident from the results of the S-N
curves. Furthermore, for the regime of applied stresses below yield stress, stage III occurs at
approximately 80 % for 2198, while below 70 % for the conventional 2024 alloy.

5.2 Energy absorbed per fatigue cycle


Absorbed energy W in kJ/m3 per fatigue cycle was evaluated by integrating the hysteresis fatigue loop of
maximum applied fatigue stress strain, as shown in Figure 7; integration has been performed per
different stages of fatigue lifetime for the same constant amplitude loading till fracture as well as for
different applied stress levels. Calculation results can be seen in Figure 10(a) and (b), for AA2024-T3 and
AA2198-T351, respectively. The diagrams containing the experimental values have as horizontal axis the
number of fatigue cycles (logarithmic scale) and as vertical axis the absorbed energy per fatigue cycle
(linear scale).
Two different regimes can be distinguished in Figure 10(a); the first deals with high applied
stresses (> Rp = 391 MPa) that corresponds to more-or-less to low cycle fatigue conditions as Nf < 104
cycles. Evaluated absorbed energy values are continuously decreasing with increasing fatigue life, e.g. for
max = 480 MPa. Almost total decrease of 40 % between the first and last cycle is very short in duration
and lasts approximately 10 to 15 % of total fatigue life. Of course this decreasing is a matter of hardening
that takes place during fatigue and is another hardening sign besides stiffness increase for the same
fatigue lifetime.
9

International Journal of Fatigue 56 (2013) 95105


However, such a decrease is also evident for the regime with constant amplitude fatigue loading
below yield stress. Again, hardening and W decrease can also be noticed; nevertheless, this decrease
seems to take lower values than the previous regime, ranging from 20 to 40 %. As expected, the lower the
fatigue applied stress, the lower the absorbed energy W of hysteresis loop is. When reducing even more
the applied fatigue stress that is far below fatigue endurance limit, then the absorbed energy values are
further reduced and their calculation accuracy might be problematic due to measuring scatter. Summingup, it is evident that hardening has been noticed for all fatigue stresses; W decrease from the initial fatigue
cycles up to a few cycles before fracture is highly dependant upon applied fatigue stress. Nevertheless,
hardening takes place at stresses even below yield stress or even fatigue endurance limit.
Figure 10(b) gives the AA2198-T351 respective calculation results; higher absorbed energy values
were calculated (ranging from 2 to 21 kJ/m3) when compared to the AA2024-T3 values. This was
expected, since overall higher tensile ductility values were calculated for the innovative alloy, Figure 3
and Table 2. Hardening with increasing fatigue life is also evident since W is decreasing, especially for
the high applied stresses regime. For the regime that max > 80 % of tensile strength Rm, W decrease was
calculated to be approximately from 39 % till 45 % of the respective initial values. For applied stresses
above and very close to yield stress (~ 250 MPa), W seems to slowly decreasing (almost 25 % total
decrease for max = 285 MPa) or not essentially decreasing at all (max = 250 MPa). This is a clear
advantage of the innovative alloy, since for fatigue applied stresses around yield stress no essential
decrease of absorbed energy per hysteresis loop has been noticed. Notice that for 2024-T3 and for the
applied stresses regime below yield stress, it showed average 35 % decrease that is evidence of changes
of the materials microstructure to support the hardening effect. However, this is not the case for the
innovative alloy and therefore it is proven to be very stable and reliable at the stress regime of and below
yield stress and in conjunction to the results of the specific S-N curves of Figure 6 proves its superiority
in the above stress regimes.
In order for a material to be used in aerostructures, it is of imperative importance for high fatigue
damage tolerance characteristics. To a lesser extent, it is essential for a material to absorb high energy
during its various mechanical loadings and without degrading its original mechanical properties. To this
end, it has been discussed in this work that AA2198-T351 has a clear advantage over AA2024-T3
regarding its fatigue mechanical performance at the regime below yield stress; it was demonstrated that W
is not essentially decreasing over fatigue life of a specimen at the referred stress levels. The authors will
also try to demonstrate the overall capacity of the investigated alloys to absorb mechanical energy under
constant amplitude fatigue loadings. The experimental curves of absorbed energy per hysteresis loop were
integrated for the complete fatigue lifetime of every specimen and as a function of maximum applied
fatigue stress. The calculated results of the integration were marked as Wall in the diagram of Figure 11.
For comparison purposes, the fatigue stress was normalized to the ultimate tensile strength (max / Rm).
10

International Journal of Fatigue 56 (2013) 95105


The integration of energies per fatigue cycle of 2024-T3 for high stresses (> 90 % of Rm) gives identical
values of energy with the tensile strain energy density value (~ 86 MJ/m3) that is marked in the diagram.
With decreasing applied fatigue stress the integration of the energy of fatigue hysteresis loops gives
continuously increasing values (exponential curve fit). Same curve fitting can also be noticed for
AA2198-T351 case; higher energy values were calculated for every applied stress level of Figure 11. It
can be seen that the integration of energies for high stresses gives slightly higher values than the
respective quasi-static value and this requires investigation in conjunction with advanced metallographic
analyses to point out the microstructural phenomena taking place during this mechanical testing.
Nevertheless, it is quite clear that the innovative alloy requires by a multiplication factor of 2 to 3 higher
total energy to fracture under constant amplitude fatigue for the same normalized stress, which is again a
parameter showing the supremacy of the innovative alloy over AA2024-T3.

5.3 Cyclic stress - strain diagrams


The experimental results of stress - strain values from the fatigue tests were used for the construction of
the cyclic stress - strain (CSS) curves. Values of the pairs of maximum applied fatigue stress (max) and
maximum nominal strain (max) at specific percentages of fatigue lifetime were taken by different constant
amplitude tests to construct a CSS curve. Figure 12 shows the cyclic stress - strain curves of AA2024-T3
at three different fatigue lifetime percentages, namely 10, 50 and 80 % of total fatigue lifetime. In the
diagram, both the quasi-static tensile curve as well as CSS curves from the open literature can also be
seen [32],[33].
For the case of applied maximum stress up to yield stress Rp, no essential differences have been
noticed between CSS curves and the respective quasi-static curve. The differences can be noticed when
the specimens were loaded at maximum stresses beyond yield stress that correspond to high and low
cycle regimes. It is obvious that the higher the investigated percentage of materials fatigue life, the more
the CSS curve tends to swift to the left of the diagram. This means that the materials flow curve is
altering and the material is toughened. This was also confirmed by the lowest absorbed energy per fatigue
loop that was discussed in a previous section.

5.4 Work hardening exponent


The CSS curves showed that AA2024-T3 work harden with increasing fatigue life. Figure 13 shows the
results of the linear part (stage II) work hardening exponent n of the investigated alloys that were
calculated from respective CSS curves. For the case of AA2024-T3, work hardening n was reduced by
11

International Journal of Fatigue 56 (2013) 95105


18.63 % after consuming only 10 % of alloys fatigue life. However, after consuming the rest 40 % of
fatigue life (from 10 to 50 %) it was further decreased by 3.22 %. The same applied also for the regime
from 50 to 80 % that only a small decrease (2.81 %) in work hardening exponent was noticed. Taking into
account the previous experimental results, it was expected that most of hardening takes place during the
earlier stages of AA2024-T3 fatigue life.
For the case of Al-Li alloy, it was demonstrated that in the first 10 % of fatigue life hardening is
clearly evident and thus reducing by almost 5 % the work hardening exponent. Please notice that the
calculation of work hardening exponent had been performed by respective CSS curves; these curves were
constructed by employing experimental data (stress - strain) values from different constant amplitude
fatigue tests. As it was shown that different hardening percentage took place per different applied max,
construction of this curve allows the calculation of an average curve with different hardening percentages
(average hardening). Hence, it can be used for comparison between the different fatigue stages of the
same alloy, e.g. AA2198 that showed no essential average hardening at stages below 20 % of fatigue
lifetime.

6. Fractography
Specific specimens that are marked in the S-N fatigue curve have been examined in a scanning electron
microscope of Sheffield Hallam University and the fractographic results will be presented beneath.
Fracture surface of three different specimens were examined that correspond to different fatigue regimes;
(a) specimen No 5 with max = 285 MPa and fatigue life ~ 1.5 x 105 cycles and (b) specimen No 4 with
max = 275 MPa and fatigue life ~ 2 x 105 cycles that both correspond to high cycle fatigue regime and
finally (c) specimen No 11 with max = 250 MPa and fatigue life ~ 8 x 105 cycles that corresponds in the
transition regime from HCF to fatigue endurance limit regime. The fatigue and fracture mechanisms for
these dissimilar specimens will be discussed.
Fractographic examination of the first specimen (sp.05) revealed edge crack initiation sites formed
leading to crack growth, Figure 14(a) and (b). The direction of the crack growth is also indicated on the
image. The surface looks flat and transgranular fractures appearing with multiple layers of macro
striations formed along the crack growth directions. Increasing levels of plasticity are causing the
fractures to initiate and grow at this point.
Figure 15(a), taking in the middle section of the fractured second specimen (sp.04), showed
different fracture characteristics, such as the formation of a shear lip. Shear lips imply that the fatigue
crack growth occurs in a mixed mode, I + II [34]. A simple mechanical explanation is that the process is
initiated by a situation of plane stress at the specimen surface, which leads to maximum shear stresses on
12

International Journal of Fatigue 56 (2013) 95105


planes inclined at 45 to the specimen surface. The sample's geometry, thickness and width have been
reported to have influenced the formation of slanted fracture, observed by the formation of shear lips [34].
A mixed fracture mode consisting of transgranular and intergranular facets was observed in the slanted
areas. The transitions can be associated with the start of shear lip growth and the transition of I and II
mode of fracture with the completion of it, i.e. the whole thickness has become slanted. The slanted
regions indicate that Kmax has increased dramatically, where cracks are linked and a single crack is formed
at fracture mode I and II. Figure 15(b) shows a ductile fracture feature resulting from void growth and
coalescence. Large dimples are present at the grain facets and a number of smaller dimples are aligned at
intergranular facets.
Fractographic examination of the third specimen (sp.11), revealed in Figure 16(a) the transition of
the smooth to rough sites where again large plastic zones formed leading to damage in the form of deep
cracks; dimples were also formed at the smooth fracture site. Figure 16(b) shows that the coalescence to
the stage II crack seems to be delayed; therefore the surface has a higher roughness, which has a lower
stress intensity state due to the lower stress level, thus lasting longer. The striations and the deformation
looks quite complex and this can be attributed not only to the fatigue mechanisms but also to the
microstructural strengthening mechanisms e.g. precipitation hardening. AA2198-T351 is hardened by
aging; previous work suggested that in such alloys, the main source of voids is the disintegration between
FeSi large inclusions, intermetallic dispersoids (Al3Zr), fine hardening precipitates (Al2CuLi, Al3Li and
Al2Cu) and matrix material [19],[35]. During mechanical loading, voids nucleate at the fine precipitates
and grow, while new voids will be generated by further disintegration of precipitates from the matrix, and
by breaking of precipitates. It appears that many multiple cracks are also initiated at grain boundaries.
Therefore it is likely to be more difficult to from stage II cracks at this mid-stage of the failure process.

13

International Journal of Fatigue 56 (2013) 95105

7. Conclusions
1. The fatigue endurance limit under constant amplitude loading and stress ratio R = 0.1 is almost 40
% below the AA2024 yield stress, while for 2198-T351 is only 8 % lower than the respective yield
stress.
2. Aluminum alloys 2024-T3 and 2198-T351 seems to exhibit fatigue endurance limit at the magnitude
of 240 5 MPa with constant amplitude fatigue loading. When taking into account density ,
AA2198 is superior to AA2024 in high cycle fatigue and fatigue endurance limit regimes.
3. It has been observed that hardening occurs in both alloys as the dynamic stiffness increases in the
first stages of fatigue life, highly depending upon applied maximum stress.
4. A decrease in dynamic stiffness over the fatigue life associated with the transition of fatigue stage II
to III was observed. The decrease point is dependant upon yield stress and fatigue maximum stress;
AA2198 showed a linear trend of these stages transition with applied fatigue stress. The improved
fatigue performance of AA2024 is evident at high stresses; stage III occurs at 60 % of fatigue life
time on the contrary to AA2198 that occurs at 20 %, when both stressed at 90 % of ultimate tensile
strength.
5. Hardening is noticed to both alloys by the decrease of the absorbed energy per hysteresis loop.
AA2198 proved to be very stable and reliable at the stress regime around yield stress where no
essential hardening was noticed and in conjunction to the results of the specific S-N curves proves
its superiority in the HCF and endurance limit regimes.
6. For the same normalized applied stresses, AA2198 was observed to absorb 2 to 3 times more energy
to fracture than 2024 thus proving damage tolerant capabilities.

Acknowledgements
The authors would like to acknowledge the Advanced Materials Laboratory of Hellenic Aerospace
Industry for providing the investigated materials and especially the Head of the Research & Development
Department Dr. Zaira Marioli-Riga, for fruitful discussions about the manuscript.

14

International Journal of Fatigue 56 (2013) 95105

References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]

[13]

[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]

Lavernia EJ, Grant NJ, Aluminium-lithium alloys, J Mater Sci 1987; 22: 1521-9.
Rioja RJ, Liu J, Evolution of Al-Li Base Products for Aerospace and Space Applications, Metall Mater Trans A 2012;
43A: 3325-37.
Heinz A, Haszler A, Keidel C, Moldenhauer S, Benedictus R, Miller WS, Recent development in aluminium alloys
for aerospace applications, Mater Sci Eng A 2000; A280: 102-7.
Steuwer A, Dumont M, Altenkirch J, Birosca S, Deschamps A, Prangnell PB, Withers PJ, A combined approach to
microstructure mapping of an AlLi AA2199 friction stir weld, Acta Mater 2011; 59: 3002-11.
Li H, Tanga Y, Zenga Z, Zhenga Z, Zhenga F, Effect of ageing time on strength and microstructures of an Al-Cu-LiZn-Mg-Mn-Zr alloy, Mater Sci Eng A 2008; 498: 314-20.
Yoshimura R, Konno TJ, Abe E, Hiraga K, Transmission electron microscopy study of the evolution of precipitates in
aged Al-Li-Cu alloys: the and T1 phases, Acta Mater 2003; 51: 4251-66.
Cassada WA, Shiflet GJ, Starke EA Jr, The Effect of Plastic Deformation on AI2CuLi (T1) Precipitation, Met Mat
Trans A 1991; 22A: 299-306.
Kumar KS, Brown SA, Pickens JR, Microstructural evolution during aging of an Al-Cu-Li-Ag-Mg-Zr alloy, Acta
Mater 1996; 44: 1899-915.
Warner T, Recently-developed aluminium solutions for aerospace applications, Mater Sci Forum 2006; 519-21: 127178.
da Costa Teixeira J, Cram DG, Bourgeois L, Bastow TJ, Hill AJ, Hutchinson CR, On the strengthening response of
aluminum alloys containing shear-resistant plate-shaped precipitates, Acta Mater 2008; 56: 6109-22.
Decreus B, Deschamps A, De Geuser F, Donnadieu P, Sigli C, Weyland M, The influence of Cu/Li ratio on
precipitation in Al-Cu-Li-x alloys, Acta Mater 2013; 61: 2207-18.
Rao JC, Payton EJ, Somsen C, Neuking K, Eggeler G, Kostka A, Dos Santos JF, Where Does the Lithium Go? - A
Study of the Precipitates in the Stir Zone of a Friction Stir Weld in a Li-containing 2xxx Series Al Alloy, Adv Eng
Mat 2010; 12: 298-303.
Le Jolu T, Morgeneyer TF, Gourgues-Lorenzon AF, Effect of friction stir weld defects on fatigue lifetime of an AlCu-Li alloy (AA-2198) 18th European Conference on Fracture: Fracture of Materials and Structures from Micro to
Macro Scale, ECF 2010; Dresden; 30 August 2010 through 3 September 2010.
Cavaliere P, Cabibbo M, Panella F, Squillace A, 2198 Al-Li plates joined by Friction Stir Welding: Mechanical and
microstructural behavior, Mater Des 2009; 30: 3622-31.
Bitondo C, Prisco U, Squillace A, Giorleo G, Buonadonna P, Dionoro G, Campanile G, Friction stir welding of
AA2198-T3 butt joints for aeronautical applications, Int J Mater Form 2010; 3: 1079-82.
Astarita A, Squillace A, Scala A, Prisco A, On the Critical Technological Issues of Friction Stir Welding T-Joints of
Dissimilar Aluminum Alloys, J Mater Eng Perf 2012; 21: 1763-71.
Le Jolu T, Morgeneyer TF, Gourgues-Lorenzon AF, Effect of joint line remnant on fatigue lifetime of friction stir
welded Al-Cu-Li alloy, Sci Tech Weld Join 2010; 15: 694-8.
Chen J, Madi Y, Morgeneyer TF, Besson J, Plastic flow and ductile rupture of a 2198 Al-Cu-Li aluminum alloy,
Comp Mater Sci 2011; 50: 1365-71.
Steglich D, Wafai H, Besson J, Interaction between anisotropic plastic deformation and damage evolution in Al 2198
sheet metal, Eng Fract Mech 2010; 77: 3501-18.
Steglich D, Wafai H, Besson J, Anisotropic Deformation and Damage in Aluminium 2198 T8 Sheets International
Journal of Damage Mechanics 2010 19: 131-52.
Vogelesang LB, Vlot A, Development of fibre metal laminates for advanced aerospace structures, J Mater Proc Tech
2000; 103: 1-5.
Merati A, A study of nucleation and fatigue behavior of an aerospace aluminum alloy 2024-T3, Int J Fatigue 2005;
27: 33-44.
Jurcevic R, DuQuesnay DL, Topper TH, Pompetzki MA, Fatigue damage accumulation in 2024-T351 aluminium
subjected to periodic reversed overloads, Int J Fatigue 1990; 12: 259-66.
Khan S, Vyshnevskyy A, Mosler J, Low cycle lifetime assessment of Al2024 alloy, Int J Fatigue 2010; 32: 1270-7.
Kim SH, Kim KS, Kim SK, Yoon YO, Cho KS, Lee K, Microstructure and mechanical properties of Eco-2024-T3
aluminum alloy, Adv Mater Res 2013; 602-4: 623-6.
Ngiau C, Kujawski D, Sequence effects of small amplitude cycles on fatigue crack initiation and propagation in 2024T351 aluminum, Int J Fatigue 2001; 23: 807-15.
Rodopoulos CA, Choi JH, De los Rios ER, Yates JR, Stress ratio and the fatigue damage map - Part II: The 2024T351 aluminium alloy, Int J Fatigue 2004; 26: 747-52.
Ludian T, Wagner L, Effect of age-hardening conditions on high-cycle fatigue performance of mechanically surface
treated Al 2024, Mater Sci Eng A 2007; 468-70: 210-3.

15

International Journal of Fatigue 56 (2013) 95105


[29]
[30]
[31]
[32]
[33]
[34]
[35]

Bray GH, Glazov M, Rioja RJ, Li D, Gangloff RP, Effect of artificial aging on the fatigue crack propagation
resistance of 2000 series aluminum alloys, Int J Fatigue 2001; 23: 265-76.
Warner JS, Gangloff RP, Alloy induced inhibition of fatigue crack growth in age-hardenable Al-Cu Alloys, Int J
Fatigue 2012; 42: 35-44.
N. Dowling. Mechanical Behavior of Materials. Prentice Hall, New Jersey, USA, 2nd edition, 1993.
Mitchell MR, Fundamentals of Modern Fatigue Analysis for Design, ASM Handbook, Volume 19: Fatigue and
Fracture, Ohaio, United States of America, 1996, p. 553-616.
Christ HJ, Cyclic Stress-Strain Response and Microstructure, ASM Handbook, Volume 19: Fatigue and Fracture,
Ohaio, United States of America, 1996, p. 171-234.
Zuidema J, Veer F, Van Kranenburg C, Shear lips on fatigue fracture surfaces of aluminum alloys, Fat Fract Eng Mat
Struct 2005; 28: 159-67.
Robson JD, Prangnell PB, Dispersoid precipitation and process modelling in zirconium containing commercial
aluminium alloys, Acta Mater 2001; 49: 599-61.

16

International Journal of Fatigue 56 (2013) 95105

List of Tables
Table 1: Chemical composition of aluminum alloys 2024-T3 and 2198-T351.
2024-T3

2198-351

Si

Fe

Cu

Mn

Mg

Cr

Zn

Ti

Al

0.50

0.50

3.8-4.9

0.3-0.9

1.2-1.8

0.1

0.25

0.15

Rem.

Cu

Li

Zn

Mn

Mg

Zr

Si

Ag

Fe

Al

2.9-3.5

0.8-1.1

0.35

0.5

0.25-0.8

0.04-0.18

0.08

0.1-0.5

0.01

Rem.

Table 2: Evaluated tensile mechanical properties of 2024 and 2198 aluminum alloys and respective literature values.
Alloy
2024-3

Type / Source
Experiments: Average
Experiments: St. Deviation

2198-T351

Experiments: Average
Experiments: St. Deviation

Rp [MPa]

Rm [MPa]

Af [%]

70.291

391

500

18.88

86.28

2.932

21

20

2.35

6.80

75.461

265

384

14.90

52.77

0.30

1.07

1.767

2.16

1.41

2198-T3

Bitondo et al. [15]

275

370

15.00

2198-T3

Astarita et. al. [16]

315

375

15.00

2198-T351

Chen et al. [18]

13.00

2198-T851

Experiments: Average

11.55

56.49

0.37

1.88

14.00

Experiments: St. Deviation

Wtens [MJ/m3]

E [GPa]

2198-T851

Chen et al. [18]

2198-T8

Steglich et al. [20]

2198-T851

Cavaliere et al. [14]

72.803

324

442

432

484

1.826
-

2.25
*

1.87
+

490

530

73.000

469

510

14.00

76.700

436

490

13.70

yield strength for 0.2% plastic strain, uniform elongation, - non available

17

International Journal of Fatigue 56 (2013) 95105

List of Figures

(b)

(c)
(a)
Figure 1: (a) Test configuration showing a mounted fatigue specimen with the attached dynamic extensometer on the
grips of the servo-hydraulic Instron 100 kN testing machine and typical geometrical dimensions of the machined
specimens of the present study: (b) tensile coupon configuration according to ASTM E8 and (c) fatigue test specimen
according to ASTM E466.

500
Aluminum alloy 2024-T3
t = 3.2 mm - 3 specimens

Axial nominal stress [MPa]

400

300
Aluminum alloy 2198-T851
t = 1.6 mm - 3 specimens

Aluminum alloy 2198-T351


t = 1.6 mm - 3 specimens

200

100

0
0

10

15

20

25

Axial nominal strain [%]


Figure 2: Typical axial nominal stress - strain tensile curves of the investigated aluminum alloys.

18

International Journal of Fatigue 56 (2013) 95105

Percentage from respective mechanical


property of 2024-T3 sheet alloy

125

100

75

50

25

2024-T3, t = 3.2 mm
E = 70.291 GPa
Rp = 391 MPa
Rm = 500 MPa
Af = 18.88 %

modulus of elasticity E
yield stress Rp
tensile strength Rm
tensile ductility Af

0
2024-T3

2198-T351

2198-T851

Sheet aluminum alloy and temper


Figure 3: Comparison between the tensile mechanical properties of the investigated aluminum alloys.

Maximum applied stress max [MPa]

550
Jurcevic et al. 2024-T351 L,
R = 0.1, t = 5.08 mm
Rodopoulos et al. 2024-T351 L,
R = 0.1, t = 5.50 mm
Ngiau and Kujawski 2024-T351 L,
R = 0.1, t = 12.70 mm

500
450
400
350
300
250

test stopped
Experimental data
Aluminum alloy 2024-T3,
L direction, t = 3.2 mm, R = 0.1
Curve fitting

200
0
100

1000

10000

100000 1000000

1E7

1E8

Fatigue cycles to fracture Nf [-]


Figure 4: Experimental fatigue S-N curve of the investigated aluminum alloy 2024-3 and comparison with the
available literature.

19

International Journal of Fatigue 56 (2013) 95105

Maximum applied stress max [MPa]

400
Experimental data 2198-351,
L direction, R = 0.1, t = 1.56 mm
Curve fitting

380
360
340

specimen No 4
max = 275 MPa

320

specimen No 5
max = 285 MPa

300
specimen No 11
max = 250 MPa

280
260
240
220

Cavaliere et al. 2198-T851,


L direction, R = 0.1, t = 5 mm
Curve fitting

200
0
100

1000

test stopped

10000 100000 1000000

1E7

1E8

Fatigue cycles to fracture Nf [-]

200
Experimental data 2024-3
Curve fitting 2024-3
Experimental data 2198-351
Curve fitting 2198-351

180

[(MPa*m )/kg]

Specific maximum applied stress max/

Figure 5: Experimental fatigue S-N curve of the investigated aluminum alloy 2198-351.

160

HCF

LCF

140

Fatigue
endurance
limit

120
100
80
test stopped

0
100

1000

10000 100000 1000000

1E7

1E8

Fatigue cycles to fracture Nf [-]


Figure 6: Comparison of the modified fatigue S-N curves of the two investigated alloys by employing specific maximum
applied tensile stresses.

20

International Journal of Fatigue 56 (2013) 95105

448

Maximum applied stress max [MPa]

444

Y=A+B*X
Parameter
Value Error
A
-213,82677
38,868435
B
66,036643
9,987754

440
slope in GPa

436
432

Alloy 2024-T3, L direction


t = 3.2 mm, R = 0.1
max = 440 MPa
Fatigue hysterisis loop
Fit linear

Eunloading= dynamic
stiffness

428
424

W = integration of fatigue
hysterisis loop

0
0,0

3,87

3,88

3,89

3,90

3,91

Axial nominal strain [%]


Figure 7: Calculation of the dynamic stiffness during unloading branch and the integral of fatigue hysteresis loop.

21

International Journal of Fatigue 56 (2013) 95105

Dynamic stiftness of fatigue loop


during unloading [GPa]

80
75

max = 220 MPa

44.9 %
51 %

max = 250 MPa


max = 280 MPa

70
65

max = 300 MPa

61.2 %

max = 350 MPa

71.4 %

max = 440 MPa

89.8 %

60

max = 480 MPa

55

max = 485 MPa

50

57 %

max = 490 MPa

0
0

98 %

99 %
Rm = 500 MPa, Rp = 391 MPa
xx.x % = percentage of Rm
100 %

Alloy 2024-T3, L direction


t = 3.2 mm, R = 0.1

20

40

60

80

100

Percentage of fatigue damage D = N / Nf [%]

(a)

Dynamic stiftness of fatigue loop


during unloading [GPa]

82
80
78
76

max = 250 MPa


max = 275 MPa

75 %
72.3 %
77.6 %

max = 285 MPa


max = 295 MPa
max = 305 MPa

80.3 %

max = 315 MPa

82.9 %

74

Alloy 2198-T351, L direction


t =1.56 mm, R = 0.1

72
70

65.7 %

Rm = 402 MPa, Rp = 265 MPa


xx.x % = percentage of Rm
89.4 %

max = 340 MPa

68
66
0

max = 360 MPa

20

94.7 %

40

60

80

100

Percentage of fatigue damage D = N / N [%]

f
(b)
Figure 8: Stiffness of the fatigue specimens during unloading as a function of applied maximum stress and percentage
of Miners fatigue damage of aluminum alloys (a) 2024-T3 and (b) 2198-T351.

22

International Journal of Fatigue 56 (2013) 95105

Percentage of fatigue damage for the


initiation of stiffness decrease Dp [%]

100

2024-T3 yield stress

80

60

40
2024-T3
2198-T351
20
2198-T351
yield stress

0
0

55 60 65 70 75 80 85 90 95
Percentage of fatigue maximum applied stress
to ultimate tensile strength ( max / Rm ) [%]

100

Figure 9: Normalized applied maximum stress to ultimate tensile strength and percentage of fatigue life where the
initiation of decrease of stiffness occurs for the investigated alloys.

23

International Journal of Fatigue 56 (2013) 95105

max = 480 MPa

Absorbed energy per fatigue cycle W [kJ/m ]

14
Alloy 2024-T3, L direction
t = 3.2 mm, R = 0.1
Rp = 391 MPa, Rm = 500 MPa
xx.x% percentage of Rm

12
10
max = 440 MPa

98 %
max = 350 MPa

89.8 %

max = 300 MPa

71.4 %

max = 280 MPa


max = 245 MPa

61.2 %

max = 220 MPa

57 %

50 %
44.9 %

0
1

10

100

1000

10000 1000001000000 1E7

Number of fatigue cycles N [-]


3

Absorbed energy per fatigue cycle W [kJ/m ]

(a)

22

max = 360 MPa

Alloy 2198-T351, L direction


t = 1.56 mm, R = 0.1
Rp = 265 MPa, Rm = 384 MPa
xx.x % = percentage of Rm

20
max = 340 MPa

18
16
14

94.7 %

max = 305 MPa

12

89.4 %

10
8

80.2 %

max = 285 MPa

75 %

4
max = 250 MPa

65.7 %

0
1

10

100

1000

10000 1000001000000 1E7

Number of fatigue cycles N [-]

(b)
Figure 10: Absorbed energy per fatigue cycle versus the number of fatigue cycles of aluminum alloy (a) 2024-T3 and (b)
2198-T351.

24

International Journal of Fatigue 56 (2013) 95105

Percentage of maximum applied stress


to ultimate tensile strength ( max / Rm ) [%]

100
Aluminum alloy 2198-T351
experiments
curve fitting
tensile strain energy density

90

80

70
Aluminum alloy 2024-T3
experiments
curve fitting
tensile strain energy density

60
50
0
0

200

400

600

800

2000

Total absorbed energy by integration of hysteresis loops


3
per fatigue test Wall [MJ/m ]
Figure 11: Total absorbed energy by integration of fatigue hysteresis loops of fatigue tests up to fracture as a function
of percentage of maximum applied stress per ultimate tensile strength of investigated aluminium alloys.

Maximum fatigue axial stress max [MPa]

550
500
450
Cyclic stress - strain (CSS) curves
10 % fatigue life
50 % fatigue life
80 % fatigue life
Cyclic (literature)

400
350
300

Alloy 2024-T3, L direction


t = 3.2 mm, R= 0.1
quasi-static curve
monotonic (literature)

250
200
0
0

10

Maximum axial nominal strain max [%]


Figure 12: Typical cyclic stress - strain curves for different percentages of fatigue life of the aluminum alloy 2024-T3.

25

Linear work hardening exponent n (Stage II) [-]

International Journal of Fatigue 56 (2013) 95105

0,16
0,15
-5.41%

-6.78%

0,14

-7.54%

0,13
Alloy 2198-T351
exponential curve fit
Alloy 2024-T3
exponential curve fit
-xx.xx% decrease from intitial property

0,12
0,11
0,10

-16.38%

0,09

-17.67%

-13.51%

0,08
0

20

40

60

80

100

Percentage of fatigue lifetime Ni / Nf [%]


Figure 13: Work hardening exponent versus the percentage of Miners fatigue damage of aluminum alloy 2024-T3.
Crack growth direction

Transgranular fracture
Striations

Edge Crack initiator

(a)
(b)
Figure 14: Alloy 2198-T351: (a) SEM fractographs of specimen No 5 with max = 285 MPa and around Nf = 1.5 x 105
cycles with crack initiators and (b) transgrannular fracture and striations formed evidence of the accumulated damage.

26

International Journal of Fatigue 56 (2013) 95105


Shear Lip

Slanted fracture sides

Coalescence
microvoids

(a)
(b)
Figure 15: (a) Shear lip formed on the 2198-T351 specimen No 4 at max = 275 MPa / Nf ~ 2 x 105 cycles due to plastic
zone accumulation indicating possible local overloading, and (b) coalescence microvoids and evidence of ductility.

Crack originated
from edge surface

Ductile areaDimples formed

Transgranular Crack

Fatigue - Initiation site

Fracture zone

Fatigue fracture path

(a)
(b)
Figure 16: (a) SEM fractograph of 2198-T351 specimen No 11 at max = 250 MPa with surface cracks and smooth to
rough transition sites and (b) plastically deformed fracture surface with evidence of fatigue striations which is
attributed to higher fatigue life approximately Nf = 8 x 105.

27
View publication stats

You might also like