Sie sind auf Seite 1von 16

metals

Article
Effect of Overaging on the Cyclic Deformation
Behavior of an AA6061 Aluminum Alloy
Kun Liu * ID
, Foisal Ahmed Mirza and Xiao Grant Chen
Department of Applied Sciences, University of Québec at Chicoutimi 555, boulevard de l’Université,
Chicoutimi, QC G7H 2B1, Canada; f4mirza@ryerson.ca (F.A.M.); xgrant_chen@uqac.ca (X.G.C.)
* Correspondence: kun.liu@uqac.ca; Tel.: +1-(418)-545-5011 (ext. 7112); Fax: +1-(418)-545-5012

Received: 12 June 2018; Accepted: 4 July 2018; Published: 7 July 2018 

Abstract: The present work encompasses the effect of overaging on the strain-controlled low-cycle
fatigue (LCF) behavior of an extruded AA6061 aluminum alloy at varying strain amplitudes. During
the T7 aging treatment, the size of precipitates increased from 60 nm under T6 conditions to 220 nm
after aging for 48 h at 200 ◦ C, leading to a decrease in the monotonic tensile strength. During the
LCF tests, nearly symmetrical hysteresis loops can be observed in the mid-life cycle under all test
conditions, whereas the first-cycle hysteresis loops were moderately inflected under long-aging
conditions. With increasing aging time, the cyclic peak stresses decreased and the plastic strain
increased. Nearly ideal Masing behavior was exhibited under T6 conditions, while it was lost
under T7 overaging conditions. The cyclic stress responses were similar under all tested conditions,
involving stabilization at low strain amplitudes and softening at high strain amplitudes, with initial
hardening for the first few cycles. Compared to the T6 condition, the fatigue life increased with
increasing T7 aging time. Various LCF parameters were estimated based on the Coffin-Manson and
Basquin relationships and on the LCF experimental results. The relationship between the fatigue
life, strength, and microstructure of the investigated AA6061 aluminum alloy under various aging
conditions was discussed.

Keywords: 6061 aluminum alloys; aging treatments; precipitates; cyclic deformation; fatigue life

1. Introduction
The 6xxx-series aluminum alloys are extensively used in the aerospace and automotive industries,
owing to their ideal mechanical properties, good corrosion resistance, as well as good formability and
weldability [1–4]. There is considerable industrial interest in these alloys, as two thirds of all extruded
products are made of aluminum, and 90% of these are made of the 6xxx-series alloys [4]. AA6061 alloys
are one of the most widely used precipitation-hardening aluminum alloys in the 6xxx series [2,3,5].
Structural components of AA6061 alloys used in engineering services would require the evaluation
of their mechanical performance under cyclic loads, because these components would unavoidably
experience dynamic loading in service, which leads to the occurrence of fatigue failure [2,6,7]. Thus,
to ensure the structural integrity and durability of such engineering components, it is essential to
understand the fatigue and cyclic deformation behavior of AA6061 aluminum alloys.
Several studies were conducted to understand the fatigue behavior of aluminum alloys over
the years, especially on age-hardening aluminum alloys, such as the 2xxx and 7xxx alloys [1,2,6–16].
Haji [6] analyzed the influence of the microstructure and alloy compositions in AA2024-T6 and
AA7020-T6 aluminum alloys on their cyclic deformation behavior, and found that the degree of
compatibility and precipitation of Zn and Cu on their parent metal led to higher fatigue properties in
the AA7020-T6 alloy than in the AA2024-T6 alloy. Lapovok et al. [12] studied the fatigue behavior of
2124 aluminum alloys, processed by equal channel angular extrusion, and reported that the low-cycle

Metals 2018, 8, 528; doi:10.3390/met8070528 www.mdpi.com/journal/metals


Metals 2018, 8, 528 2 of 16

fatigue (LCF) life was improved owing to the grain refinement. However, the existing systematic
studies on the LCF behaviors of 6xxx alloys are limited [1,2,9,13]. Wong et al. [9] conducted cyclic
strain-controlled fatigue tests on an extruded 6061-T6 alloy, and Brammer et al. [1] conducted fatigue
tests on an extruded 6061-T6 alloy at a fixed 5-Hz frequency at different strain amplitudes. In both
cases, the tests were performed up to the maximum 0.6% strain amplitude, which was not sufficient
to fully reveal the LCF behavior of the AA6061 alloys. LCF studies related to high strain amplitudes
(0.8–1.2%) and performed at different frequencies for different strain amplitudes remain to be reported
in AA6061 alloys.
Heat treatment and especially aging treatment are of great significance for precipitation-hardening
aluminum alloys. However, only limited research is found to investigate the influence of various heat
treatments on LCF behavior, especially for 6xxx aluminum alloys. Adnan et al. [2] investigated the
LCF behavior of an AA6061 aluminum alloy in three heat-treatment conditions (annealing (O), T4,
and T651). It was found that the 6061-O alloy had the highest value of transition fatigue life, owing to
high ductility, but little explanation on the relationship between the LCF behavior and precipitates
was given. Nandy et al. [17] studied the LCF performance of an AA6063 alloy subjected to under-aged,
peak-aged, and overaged conditions. It was revealed that the cyclic hardening and softening properties
were quite different under the three aging conditions. The structural components of AA6061 alloys are
mostly used at the peak-aged (T6) and overaged (T7) conditions. Owing to the change in precipitate
characteristics (mainly the number density, size, and volume fraction), the T6- and T7-treated alloys
exhibit different combinations of strength and ductility, which are expected to have a significant effect
on the LCF behavior, and consequently, on the loadbearing capability and endurance of a structural
component. It is, therefore, of great technical interest to investigate the cyclic stress-strain behavior,
including the cyclic stress and strain response, hysteresis loops, and fatigue life of AA6061 alloys
during various aging treatments. Such information on cyclic strain-controlled behavior is essential
because of the potential use of the alloy in fatigue-critical and temperature-sensitive applications [2].
In the presented work, overaging treatments were performed on an extruded AA6061 aluminum
alloy to investigate the effect of various degrees of overaging on LCF behavior at various strain
amplitudes. The LCF behaviors under various overaging conditions were systematically analyzed
using hysteresis loops, cyclic stress-strain responses, and multiple fatigue parameters, and compared
to the properties under peak-aging conditions. The relationship between the precipitation, mechanical
properties, and LCF performance of an AA6061 aluminum alloy was discussed.

2. Materials and Experimental Procedure


The material under investigation was an extruded AA6061 aluminum alloy with a chemical
composition as given in Table 1. The material was received under T6 conditions as cylindrical rods.
The T6 temper designation indicates that this material was solution-treated, and was then artificially
aged at 180 ◦ C for 8 h to peak conditions. The rods were further aged at 200 ◦ C for 5, 24, and 48 h to
reach different degrees of overaging. The four aging conditions applied in the presented work are
hereafter named as “T6”, “T7-5”, “T7-24”, and “T7-48” in the text, respectively.

Table 1. Chemical composition of the extruded AA6061 alloy (wt. %).

Material Mn Si Cu Cr Mg Fe Al
6061 0.045 0.63 0.26 0.05 0.85 0.25 Balance

Sub-sized tensile and fatigue samples were machined with the loading axis parallel to the
extrusion direction. The dimensions of samples for the tensile and LCF tests are shown in Figure 1.
Prior to testing, the gauge sections of the tensile and fatigue specimens were progressively ground
along the loading direction with emery papers up to a grit number of 600 to remove residual stresses
and any machining marks. Tensile tests were performed in accordance with the ASTM E8 standard
Metals 2018, 8, 528 3 of 16
Metals 2018, 8, x FOR PEER REVIEW 3 of 15

in in
ananInstron
Instron8801 8801 servo-hydraulic testingsystem
servo-hydraulic testing system(Instron,
(Instron, Norwood,
Norwood, MA, MA,
USA) USA)
at a at a strain
strain rate
rate of
10−3−s3−1s−at1 at
of 11 ××10 room temperature. Strain-controlled (total strain) “pull/push”-type
room temperature. Strain-controlled (total strain) “pull/push”-type fatigue tests, in fatigue tests,
in accordance
accordance with the the ASTM:
ASTM:E606 E606standard,
standard,werewere conducted
conducted in in
air air
at room temperature
at room temperaturewithwith
a 25-mm
a
extensometer
25-mm extensometerusing a fullyusingcomputerized Instron 8801Instron
a fully computerized servo-hydraulic testing system
8801 servo-hydraulic operated
testing systemwith
theoperated
Bluehill with LCF3the Bluehill
software LCF3 3,
(Version software
Instron,(Version
Norwood, 3, MA,
Instron,
USA). Norwood, MA, USA). Thetest
The cyclic-deformation
cyclic-deformation
conditions consisted test of a conditions
zero-meanconsisted of aastrain
strain (i.e., zero-mean
ratio ofstrain
Rε =(i.e.,
−1, acompletely
strain ratioreversed
of Rε = strain
−1,
completely
cycle) reversedstrain
and a constant strainrate
cycle)
of 1and
× 10 −2 s−1 with
a constant strain rate of 1 ×loading
a triangular 10−2 s−1 with a triangular
waveform. loading in
As preferred
thewaveform.
ASTM: E606 As preferred
standard for in the ASTM: E606
continuous standard
cyclic forgenerally
tests, and continuous forcyclic tests, and generally
strain-rate-sensitive for
materials,
strain-rate-sensitive materials, a triangular waveform results in a constant
a triangular waveform results in a constant strain rate during the course of one cycle. The cyclic strain rate during the
course ofwas
frequency one varied
cycle. The cyclic frequency
depending was varied
on the strain depending
amplitude on the strain
to maintain amplitude
a fixed to maintain
strain rate. The cyclic
a fixed strain rate. The cyclic frequency was calculated using the
frequency was calculated using the triangular waveform, and it was varied depending on thetriangular waveform, and it was
strain
varied depending on the strain amplitude to maintain a fixed strain rate. LCF tests were performed
amplitude to maintain a fixed strain rate. LCF tests were performed at different strain amplitudes
at different strain amplitudes ranging from 0.2% to 1.2%. At least two samples were tested at each
ranging from 0.2% to 1.2%. At least two samples were tested at each level of the strain amplitude
level of the strain amplitude to confirm the results. At low strain amplitudes (e.g., 0.2%),
to confirm the results. At low strain amplitudes (e.g., 0.2%), strain-controlled tests were sustained
strain-controlled tests were sustained for 10,000 cycles before being converted to load control, with a
for 10,000 cycles before being converted to load control, with a sine cyclic waveform at a frequency
sine cyclic waveform at a frequency of 50 Hz. The fatigue life was considered as the number of cycles
of required
50 Hz. The fatigue life was considered as the number of cycles required to completely separate
to completely separate the samples.
the samples.

Figure1. 1.Dimensions
Figure Dimensions of
of samples for the
samples for the tensile
tensile test
testand
andthe
thelow-cycle
low-cyclefatigue
fatigue (LCF)
(LCF) testtest
in in
thethe
presented work.
presented work.

The
The evolution
evolution of microstructure,
of the the microstructure, particularly
particularly the intermetallics
the intermetallics and
and grains, grains,
were were
characterized
characterized using optical microscopy and scanning electron microscopy (SEM, JSM-6480
using optical microscopy and scanning electron microscopy (SEM, JSM-6480 LV, JEOL, Tokyo, Japan). LV,
JEOL, Tokyo, Japan). Transmission electron microscopy (TEM, JEM-2100, JEOL, Tokyo, Japan) was
Transmission electron microscopy (TEM, JEM-2100, JEOL, Tokyo, Japan) was applied with an operating
applied with an operating voltage of 200 kV to observe the Mg2Si precipitates under various aging
voltage of 200 kV to observe the Mg2 Si precipitates under various aging conditions. The TEM samples
conditions. The TEM samples were prepared by twin-jet electropolishing. After the LCF tests, the
were prepared by twin-jet electropolishing. After the LCF tests, the fracture surfaces of samples were
fracture surfaces of samples were examined with SEM, with the aim of identifying the various
examined with SEM, with the aim of identifying the various features involved in the fatigue initiation
features involved in the fatigue initiation and propagation mechanisms.
and propagation mechanisms.
3. Results and Discussion
3. Results and Discussion
3.1. Microstructure Evolution During Aging Treatment
3.1. Microstructure Evolution During Aging Treatment
Figure 2 shows the typical optical micrographs of the investigated 6061 alloy under T6
Figure 2 shows the typical optical micrographs of the investigated 6061 alloy under T6 and T7-48
and T7-48 conditions. Generally, two types of particles can be found in all four aging conditions: the
conditions. Generally, two types of particles can be found in all four aging conditions: the gray Al-Fe-Si
gray Al-Fe-Si intermetallics and the black undissolved Mg2Si particles, both of which are fine due to
intermetallics and the black undissolved Mg2 Sican
the extrusion process [18]. No obvious change
particles, both of which are fine due to the extrusion
be observed in the size and the volume fraction
process [18]. No obvious change can be observed in the
of the two intermetallic particles, as aging at 200 °C could not size andany
cause theremarkable
volume fraction
changesofin the two
these
intermetallic particles, ◦
intermetallic particlesas[19].
aging at 200 C could not cause any remarkable changes in these intermetallic
particles [19].
To reveal the grain size and structure, electron back-scattered diffraction (EBSD) mapping was
To revealfor
performed theallgrain
four size
agedand structure,
samples. electron
The EBSD back-scattered
results diffraction
show that the extruded(EBSD) mapping
materials was
were all
performed for all four aged samples. The EBSD results show that the extruded materials were all fully
Metals 2018, 8, 528 4 of 16

Metals 2018, 8, x FOR PEER REVIEW 4 of 15


Metals 2018, 8, x FOR PEER REVIEW 4 of 15
recrystallized with equiaxed grains, and no difference could be found under the four aging conditions.
fully recrystallized with equiaxed grains, and no difference could be found under the four aging
As an example,
fully
conditions.
the EBSD
recrystallized withmapping
As an example, equiaxed
the EBSD
images
grains, of
and
mapping
T6images
noand T7-48
differenceare shown
could
of T6 and
in Figure
be are
T7-48 found 3.
under
shown
Itthe
canfour
in Figure
be observed
3. aging
It can
thatconditions.
bethe equiaxed
observed As
thatgrains
an the are all
example, theover
equiaxed EBSD themapping
grains sample
are section
images
all over ofinT6
the sample both
andagingin conditions,
T7-48
section are shown
both with
in a similar
Figure
aging conditions, 3. with grain
It can
a
sizebe
in the range
observed of
that 60–100
the µm.
equiaxed grains are
similar grain size in the range of 60–100 μm. all over the sample section in both aging conditions, with a
similar grain size in the range of 60–100 μm.

Figure 2. Microstructure of the experimental 6061 alloy under (a) T6 and (b) T6 further aged for 48 h
Figure 2. Microstructure of the experimental 6061 alloy under (a) T6 and (b) T6 further aged for 48 h
Figure
(T7-48) 2. Microstructure of the experimental 6061 alloy under (a) T6 and (b) T6 further aged for 48 h
conditions.
(T7-48) conditions.
(T7-48) conditions.

Figure 3. Electron back-scattered diffraction (EBSD) mapping results showing the grain structure
Figure
under
Figure 3. T6
(a) Electron
(b)back-scattered
andback-scattered
3. Electron diffraction
T7-48 conditions.
diffraction (EBSD)
(EBSD) mapping
mapping results
results showing
showing thethe grain
grain structure
structure under
under (a) T6 and (b) T7-48 conditions.
(a) T6 and (b) T7-48 conditions.
Figure 4 shows bright-field TEM images of the precipitates in the aluminum matrix for the four
agingFigure 4 shows
conditions. Asbright-field
shown in TEM Figure images
4a, the of precipitates
the precipitates
wereinverythe aluminum
fine with matrix for the
an average four
length
Figure
aging 4 shows As
conditions. bright-field
shown in TEM
Figureimages
4a, theofprecipitates
the precipitates
were in thefine
very aluminum
with an matrix
average for the four
length
of 60 nm under T6 conditions. The precipitates under T7-5 conditions were similar to those under
aging
of conditions.
T6 60 nm under
conditions As
in T6 shown
conditions.
terms in morphology,
of their Figure
The 4a, the
precipitatesbutprecipitates
under were
lengthvery
T7-5 conditions
the average finesimilar
were
increased with
to 85 an to average
nm those
(Figure underlength
4b).
of 60
T6 nm under
conditions T6
in conditions.
terms of theirThe precipitates
morphology, but under
the T7-5
average conditions
length
However, as the aging time at 200 °C increased to 24 h and 48 h, a remarkable increase in the were
increased tosimilar
85 nm to those
(Figure under
4b).
T6 conditions
However,
precipitate as in
size terms
thecan agingof their
time in
be found morphology,
at comparison
200 °C increasedbut the average
to thattounder
24 h T6
and length increased
48 h, a remarkable
conditions to
(Figure 4c,d). 85 nm
increase (Figure
in the 4b).
For instance,
precipitate size can be found ◦
in comparison totothat
24 hunder T6h,conditions
However,
the averageas the aging
length time
of theatprecipitates
200 C increased
reached 220 andafter
nm 48 48 h at 200(Figure
a remarkable 4c,d).Figure
°C increase
(T7-48, For
in theinstance,
precipitate
4d). In
sizethe
canaverage thelength
be found
addition, of density
the precipitates
in comparison
number to that reached
(volume under
fraction) T6220 nm after (Figure
ofconditions
precipitates 48with
h at longer
200 °Caging
4c,d). (T7-48,
For Figure
instance,
times was 4d).
the In
average
much
addition,
lower than the number
that under density
T6 (volume
conditions. fraction)
The of
precipitation ◦
precipitates with
microstructure longer
during
length of the precipitates reached 220 nm after 48 h at 200 C (T7-48, Figure 4d). In addition, the number aging
agingtimes
for was
6061 much
alloys
lower
(Mg2Si
density than
(volume thatfraction)
precipitates)underisT6 of conditions.
well known in
precipitates Thethe
withprecipitation
literature
longer aging microstructure
[3,19–21].
times was duringlower
Comparing
much aging
the for 6061
morphology
than that alloys
and T6
under
(Mg 2
conditions.Si precipitates)
length of The the precipitates is
precipitation well known
(Figure in the
4) to those in
microstructure literature [3,19–21].
the literature,
during aging for the Comparing
dominant
6061 the morphology
alloys precipitates under theand
(Mg2 Si precipitates) T6well
is
length
and T7-5of the precipitates
conditions were (Figure
most 4) to those
likely the in the literature,β”-Mg
semi-coherent the dominant
2Si, which precipitates
is reportedunderto thestill
be T6
known in the literature [3,19–21]. Comparing the morphology and length of the precipitates (Figure 4)
and T7-5 conditions
shearable while were mostprecipitates
likely the semi-coherent β”-Mg 2Si, which is reported to be still
to those in the[17],
literature, the
themajor
dominant underunder
precipitates the T7-24
the T6 andand T7-48
T7-5conditions
conditionswere wereβ′-Mg
most2Si, likely
shearable [17], while the
which is non-coherent andmajor precipitates
non-shearable [21].under the T7-24 and T7-48 conditions were β′-Mg2Si,
the semi-coherent β”-Mg2 Si, which is reported to be still shearable [17], while the major precipitates
which is non-coherent and non-shearable [21].
under the T7-24 and T7-48 conditions were β0 -Mg2 Si, which is non-coherent and non-shearable [21].
Metals 2018, 8, 528 5 of 16
Metals 2018, 8, x FOR PEER REVIEW 5 of 15

Figure 4. Transmission electron microscopy (TEM) bright-field images of the precipitates under
Figure 4. Transmission
various electron
aging conditions. microscopy
(a)T6; (b)T7-5; (c)(TEM)
T7-24; bright-field
and (d)T7-48.images of the precipitates under various
aging conditions. (a)T6; (b)T7-5; (c) T7-24; and (d)T7-48.
3.2. Monotonic Tensile Properties under Various Aging Conditions
3.2. Monotonic Tensile Properties under Various Aging Conditions
Typical tensile properties of the investigated 6061 alloy obtained under various aging
Typical tensile
conditions properties
are listed in Tableof2.the
Asinvestigated
seen from Table 6061 2,
alloy
bothobtained under
the yield various
strength (YS)aging conditions
and ultimate
aretensile
listed strengths
in Table 2.(UTS) decreased,
As seen whereas
from Table thethe
2, both elongation increased,
yield strength withultimate
(YS) and increasing agingstrengths
tensile time.
Only small changes in the strength were observed between the T6 and T7-5
(UTS) decreased, whereas the elongation increased, with increasing aging time. Only small changes conditions, and
in between
the strengththe T7-24 and T7-48 between
were observed conditions.
theHowever,
T6 and T7-5a large difference
conditions, in between
and strength occurred
the T7-24between
and T7-48
these two However,
conditions. groups. For instance,
a large YS onlyindecreased
difference by 11 MPa
strength occurred betweenthese
between T6 (286
twoMPa) andFor
groups. T7-5 (275
instance,
YS only decreased by 11 MPa between T6 (286 MPa) and T7-5 (275 MPa), but YS dropped by 67the
MPa), but YS dropped by 67 MPa from T7-5 (275 MPa) to T7-24 (208 MPa). The evolution of MPa
mechanical
from T7-5 (275properties was (208
MPa) to T7-24 highly related
MPa). The to the precipitates
evolution under various
of the mechanical agingwas
properties conditions. As
highly related
to shown in Figure 4, the size of precipitates increased and the number density of precipitates
the precipitates under various aging conditions. As shown in Figure 4, the size of precipitates
decreased with aging time, leading to decreasing strength according to the Orowan strengthening
increased and the number density of precipitates decreased with aging time, leading to decreasing
mechanism [22]. Corresponding to the strength change, it can be observed that the precipitate sizes
strength according to the Orowan strengthening mechanism [22]. Corresponding to the strength
were similar between T6 and T7-5 (in the range of 60–90 nm), and between T7-24 and T7-48 (in the
change, it can be observed that the precipitate sizes were similar between T6 and T7-5 (in the range of
range of 200–220 nm). However, the morphology and type of the precipitates in these two groups
60–90 nm), and between T7-24 and T7-48 (in the range of 200–220 nm). However, the morphology and
changed significantly, whereby β”-Mg2Si precipitates of fine size were dominant in T6 and T7-5,
type of the
while precipitates
β’-Mg in these two groups changed significantly, whereby β”-Mg2 Si explains
2Si precipitates of coarse size were prevalent in T7-24 and T7-48, which
precipitates
the of
fine size drop
wereindominant in T6T6and T7-5, while 0
β -Mg2 Si precipitates of coarse size were prevalent in
sharp strength from to T7-24 and T7-48.
T7-24 and T7-48, which explains the sharp drop in strength from T6 to T7-24 and T7-48.
Metals 2018, 8, 528 6 of 16

Table 2. Monotonic material properties under various aging conditions (T6, and T6 further aged for
5 (T7-5), 24 (T7-24), and 48 (T7-48) hours).

Yield Ultimate n, K/MPa, Monotonic


Elongation
Condition Strength YS Strength UTS UTS/YS Strain-Hardening Strength
El (%)
(MPa) (MPa) Exponent Coefficient
T6 286 (5.2) * 319 (6.7) 1.12 14.6 (2.3) 0.10 439
T7-5 275 (4.8) 305 (6.1) 1.11 15.5 (1.5) 0.10 424
T7-24 208 (4.1) 244 (5.8) 1.17 17.9 (2.3) 0.12 365
T7-48 188 (4.5) 228 (5.2) 1.21 19.6 (2.8) 0.14 356
6061-T6 [2] 300 338 1.13 13 N/A 480
* Note: standard deviation is shown in brackets.

It is also interesting to note that the strain-hardening exponent (n) increased and the strength
coefficient (K) decreased with increasing aging time. Similar results were reported for the AA6061-T6
alloy by Adnan et al. [2], and for the AA2024-T6 and AA7020-T6 alloys by Haji [6]. It was reported
that the strain-hardening exponent has a major influence on the forming operation, which controls the
amount of uniform plastic strain in the material before strain localization or necking [23]. Therefore,
with an increasing strain-hardening exponent, more plastic deformation occurs after the yield point, but
before the necking; consequently, the ductility of materials is improved, exemplified by the elongation
shown in Table 2. On the other hand, the strength coefficient (K) represents the real stress when the real
strain is 1, and it is greatly correlated to the treatment conditions of a material [24]. In the presented
work, the real stress decreased owing to the coarsening of precipitates with increasing aging time and
temperature, resulting in a decreasing K value. Compared to the mechanical properties under the
same T6 conditions in the literature [2], the experimental alloy had similar mechanical properties.

3.3. LCF Fatigue Behavior under Various Aging Conditions

3.3.1. Hysteresis Loops


In order to show the LCF behavior trends of the investigated 6061 alloy, the strain amplitudes
of 0.4%, 0.8%, and 1.2% were selected to represent the LCF results at low, middle, and high strain
amplitudes. The hysteresis loops of the first cycle and the mid-life stable cycle under various aging
conditions are illustrated in Figure 5. Nearly symmetrical hysteresis loops could be observed under
all aging conditions at all strain amplitudes, which were similar to those of face-centered cubic
(FCC) metals (e.g., Al, Cu, and Ni) as a result of the dislocation-slip-dominated deformation in most
materials [25].
However, despite the similarity, there were differences under various aging conditions. Firstly,
the shape of loops with longer aging time at T7 became inflected, especially at higher strain amplitudes.
As indicated by the arrows in Figure 5a,c,e, the loops were noticeably flattened after T7-48, which was
also reported in the work of Hidayetoglu et al. [8]. The inflection was only present during the first
few cycles, and disappeared gradually as cycling proceeded. As shown in Figure 5b,d,f, no obvious
inflection could be observed during the mid-life cycle. The inflection for the first few cycles with
longer aging time can be attributed to the different precipitates formed during the aging treatment [8].
For the T6 and T7-5 conditions, as shown in Figure 4, the precipitates could be sheared [17] during
the “to-and-fro” motion of dislocations during LCF, leading to the “normal” hysteresis loop without
inflection. Nevertheless, the precipitates were non-shearable under the T7-24 and T7-48 conditions.
These non-shearable precipitates pinned the movement of dislocations by forming a dislocation loop,
resulting in a population of mobile dislocations around each precipitate when the forward stress was
first applied. These dislocation loops can provide a limited amount of strain in the reverse direction
at relatively low stress. Only when they are exhausted will the reverse stress increase to a value
approaching that of the forward stress in order to continue reverse straining, thereby causing the
inflection in the first few cycles. With increasing cycles, the contribution of dislocation loops diminishes,
and then, the inflection disappears [8].
Metals 2018, 8, 528 7 of 16
Metals 2018, 8, x FOR PEER REVIEW 7 of 15

Figure
Figure 5. First
5. First andand mid-life
mid-life cyclehysteresis
cycle hysteresis loops
loops under
undervarious
variousaging conditions
aging at given
conditions strainstrain
at given
amplitudes of (a,b) 0.4%, (c,d) 0.8%, and (e,f) 1.2%.
amplitudes of (a,b) 0.4%, (c,d) 0.8%, and (e,f) 1.2%.

Secondly, the maximum/minimum peak stresses decreased with increasing T7 aging time at all
Secondly, the maximum/minimum
strain amplitudes. In the first cycle, the peak stresses decreased
maximum/minimum peak with increasing
stresses T7 aging
were similar time
between T6 at all
strainand
amplitudes. In the
T7-5 to those first the
between cycle, theand
T7-24 maximum/minimum
T7-48 conditions, butpeak
they stresses
decreasedwere
fromsimilar
T6/T7-5between
to
T7-24/T7-48, which was similar to the evolution of the monotonic mechanical properties
T6 and T7-5 to those between the T7-24 and T7-48 conditions, but they decreased from T6/T7-5 to (Table 2).
In the mid-life
T7-24/T7-48, whichcycle,
was the difference
similar to theinevolution
peak stresses between
of the each aging
monotonic conditionproperties
mechanical became larger,
(Table 2).
indicating the varying cyclic stress and strain response under various aging conditions. Table 3 lists
In the mid-life cycle, the difference in peak stresses between each aging condition became larger,
the differences in peak stresses and plastic strain between T6 and various T7 conditions during the
indicating the varying cyclic stress and strain response under various aging conditions. Table 3 lists
first and mid-life cycles. It was also noted that the plastic strain generally increased with increasing
the differences in peak stresses and plastic strain between T6 and various T7 conditions during the
first and mid-life cycles. It was also noted that the plastic strain generally increased with increasing
aging time. Taking the strain amplitude of 0.8% as an example, as shown in Table 3, the difference
in maximum peak stress between T6 and T7-5 was 14 MPa, but it increased between T6 and T7-24
Metals 2018, 8, 528 8 of 16

(43 MPa) during the first cycle. During the mid-life cycle, the differences became even larger, e.g.,
Metals 2018, 8, x FOR PEER REVIEW 8 of 15
it was 19 MPa between T6 and T7-5, and 53 MPa between T6 and T7-24. For the changes in plastic
strainaging
relative
time.toTaking
T6, it the
increased from 0.08ofat
strain amplitude T7-5
0.8% as to
an 0.10 at T7-24,
example, and in
as shown further increased
Table 3, to 0.13 at
the difference
T7-48induring
maximum the peak
first stress
cycle,between
whereas T6itand
increased
T7-5 wasfrom 0.05but
14 MPa, to 0.07, and further
it increased betweentoT60.11
andduring
T7-24 the
mid-life cycle.during
(43 MPa) The decrease in cyclic
the first cycle. stresses
During with increasing
the mid-life T7 aging became
cycle, the differences time can be larger,
even attributed
e.g., itto the
was 19 MPaofbetween
transformation T6 andas
precipitates, T7-5, and in
shown 53 Figure
MPa between T6 and
4, whereas theT7-24. For the
changes changes
with in such
cycles, plastic
as the
strain relative
differences in peaktostresses
T6, it increased fromstrain
and plastic 0.08 atduring
T7-5 tothe
0.10first
at T7-24,
cycle and
andfurther increased
mid-life to 0.13
cycle were at
probably
due toT7-48
theirduring
differentthe first
cycliccycle, whereas it increased from 0.05 to 0.07, and further to 0.11 during the
responses.
mid-life cycle. The decrease in cyclic stresses with increasing T7 aging time can be attributed to the
transformation
Table of precipitates,
3. Differences as shown
in peak stresses in Figure
and plastic 4, between
strain whereas T6
theand
changes with
various T7cycles, suchduring
conditions as the
differences in peak stresses and plastic strain during the first cycle and mid-life cycle were probably
the first and mid-life cycles.
due to their different cyclic responses.
First Cycle Mid-Life Cycle
Table
Strain3. Differences in peak stresses and plastic strain between T6 and various T7 conditions during
the first (%)
Amplitude Condition
and mid-life cycles. ∆σmax ∆σmin ∆σmax ∆σmin
∆ε (%)
p ∆ε (%) p
(MPa) (MPa) (MPa) (MPa)
Strain First Cycle Mid-Life Cycle
T7-5 −8.91 0.67 0.26 −21.27 −12.58 0.10
Amplitude Condition Δσmax Δσmin Δσmax Δσmin
0.4 T7-24 −36.97 30.03 Δε− 0.19
p (%) −61.39 27.60 Δε0.09
p (%)
(%) (MPa) (MPa) (MPa) (MPa)
T7-48 −65.82 50.60 0.12 −83.76 46.48 0.11
T7-5 −8.91 0.67 0.26 −21.27 −12.58 0.10
0.4 T7-5
T7-24 −14.15
−36.97 18.70
30.03 0.08
−0.19 −32.39
−61.39 34.79
27.60 0.05
0.09
0.8 T7-24
T7-48 − 42.67
−65.82 53.12
50.60 0.10
0.12 − 64.08
−83.76 67.48
46.48 0.07
0.11
T7-48
T7-5 −50.68
−14.15 66.07
18.70 0.13
0.08 −90.90
−32.39 90.84
34.79 0.11
0.05
0.8 T7-5
T7-24 −2.14
−42.67 9.79
53.12 −0.10
0.10 −21.54
−64.08 21.93
67.48 − 0.11
0.07
1.2 T7-24
T7-48 −40.87
−50.68 43.40
66.07 −0.01
0.13 −56.92
−90.90 59.71
90.84 0.05
0.11
T7-48
T7-5 −63.02
−2.14 65.37
9.79 0.04
−0.10 −83.15
−21.54 79.09
21.93 0.09
−0.11
1.2 T7-24 −40.87 43.40 −0.01 −56.92 59.71 0.05
T7-48 −63.02 65.37 0.04 −83.15 79.09 0.09
Figure 6 presents the stable stress-strain hysteresis loops in the mid-life cycle at various strain
amplitudes, plotted
Figure in relative
6 presents coordinates
the stable stress-straincorresponding
hysteresis loopsto
in athetranslation of the
mid-life cycle loop ascending
at various strain
branches in suchplotted
amplitudes, a way that their tips
in relative coincided
coordinates at the positions
corresponding of the loadofreversal
to a translation the loopinascending
compression.
T6 andbranches
T7-48 in suchselected
were a way that totheir tips coincided
represent the twoat extreme
the positions of the loadThe
conditions. reversal in compression.
difference between the
T6 and T7-48 were selected to represent the two extreme conditions. The difference
two conditions is obvious in terms of Masing behavior. Masing behavior is defined by comparing between the two the
shapes of hysteresis loops with the cyclic stress-strain curves drawn in the mode mentionedthe
conditions is obvious in terms of Masing behavior. Masing behavior is defined by comparing above.
If theshapes
shapeofofhysteresis
the looploops
matches with the cyclic stress-strain curves drawn in the mode mentioned above.
with the cyclic stress-strain curve, Masing behavior is considered
If the shape of the loop matches with the cyclic stress-strain curve, Masing behavior is considered to
to apply [26]. As shown in Figure 6a, the ascending branches of the loops obtained for T6 were
apply [26]. As shown in Figure 6a, the ascending branches of the loops obtained for T6 were almost
almost coincident, exhibiting nearly ideal Masing behavior [27]. On the contrary, the alloy under T7-48
coincident, exhibiting nearly ideal Masing behavior [27]. On the contrary, the alloy under T7-48
conditions clearly
conditions deviated
clearly deviated from
fromMasing
Masingbehavior (Figure6b),
behavior (Figure 6b),ininwhich
which thethe curves
curves were
were separated
separated
with increasing totaltotal
with increasing strain amplitude,
strain amplitude, especially
especiallyinin the positionsindicated
the positions indicated with
with a blue
a blue circle.
circle.

Figure 6. Superimposed stress-strain hysteresis loops with matched lower tips in the mid-life cycle at
Figure 6. Superimposed stress-strain hysteresis loops with matched lower tips in the mid-life cycle at
various total strain amplitudes: (a) T6 and (b) T7-48 conditions.
various total strain amplitudes: (a) T6 and (b) T7-48 conditions.
Metals 2018, 8, 528 9 of 16
Metals 2018, 8, x FOR PEER REVIEW 9 of 15

Masingbehavior
Masing behavioris is reported
reported to present
to be be present
if theifdislocation-dislocation
the dislocation-dislocation interaction
interaction plays aplays
morea
more important
important rolethe
role than than the dislocation–precipitate
dislocation-precipitate interaction
interaction duringduring the plastic
the plastic deformation
deformation of anyof
any multiphase material [27]. Therefore, it can be inferred that the interaction between dislocations
multiphase material [27]. Therefore, it can be inferred that the interaction between dislocations was
was favored
favored under
under T6 T6 conditions,
conditions, with behavior
with Masing Masing behavior
owing to owing to the precipitates.
the shearable shearable precipitates.
However,
However,
the the dislocation–precipitate
dislocation–precipitate interactioninteraction playedrole
played a major a major
underrole under
T7-48 T7-48 conditions,
conditions, owing toowing
the
to the non-shearable precipitates after longer T7
non-shearable precipitates after longer T7 aging times. aging times.

3.3.2.Cyclic
3.3.2. CyclicStress
Stressand
andStrain
StrainResponses
Responses
Theevolution
The evolution of stress
of stress amplitudes
amplitudes with respect
with respect to the of
to the number number ofvarious
cycles at cycles strain
at various strain
amplitudes
amplitudes
(0.4%, (0.4%,
0.8%, and 0.8%,
1.2%) and 1.2%)
is shown is shown
in Figure in various
7, under Figure 7, under
aging various aging
conditions, with aconditions, with a
semi-logarithmic
semi-logarithmic
scale scale along the X-axis.
along the X-axis.

Figure7.7.Stress
Figure Stressamplitude
amplitudevs.vs.the
thenumber
numberofofcycles
cyclesunder
undervarious
variousaging
agingconditions
conditionstested
testedatatvarious
various
strain amplitudes of (a) 0.4%, (b) 0.8%, and (c) 1.2%.
strain amplitudes of (a) 0.4%, (b) 0.8%, and (c) 1.2%.

It was observed from Figure 7 that the stress amplitude augmented, whereas the fatigue life
It was observed from Figure 7 that the stress amplitude augmented, whereas the fatigue life
decreased with increasing strain amplitude. It can be found that the stress amplitudes of T6 and
decreased with increasing strain amplitude. It can be found that the stress amplitudes of T6 and T7-5
T7-5 were always higher than those of T7-24 and T7-48 at all strain amplitudes tested. In addition, a
were always higher than those of T7-24 and T7-48 at all strain amplitudes tested. In addition, a similar
similar cyclic response was observed for all aging conditions, although it was different at applied
cyclic response was observed for all aging conditions, although it was different at applied strain
strain amplitudes. At the low strain amplitude of 0.4% (Figure 7a), cyclic stabilization occurred
amplitudes. At the low strain amplitude of 0.4% (Figure 7a), cyclic stabilization occurred under all
under all aging conditions. At the middle strain amplitude of 0.8% (Figure 7b), initial hardening can
aging conditions. At the middle strain amplitude of 0.8% (Figure 7b), initial hardening can be observed
be observed during the first few cycles (~10), followed by stabilization from T6 to T7-48. When the
during the first few cycles (~10), followed by stabilization from T6 to T7-48. When the higher strain
higher strain amplitude was applied (1.2% in Figure 7c), hardening appeared during the first few
amplitude was applied (1.2% in Figure 7c), hardening appeared during the first few cycles, followed by
cycles, followed by softening until failure for all aging conditions. Because the hardening only
softening until failure for all aging conditions. Because the hardening only happened during the first
happened during the first few cycles, the cyclic response for all aging conditions can be described as
few cycles, the cyclic response for all aging conditions can be described as stabilization at lower stain
stabilization at lower stain amplitudes, and softening at higher strain amplitudes. However, minor
amplitudes, and softening at higher strain amplitudes. However, minor differences can be found in
differences can be found in the hardening or softening rates under various aging conditions. As
the hardening or softening rates under various aging conditions. As shown in Figure 7c, the hardening
shown in Figure 7c, the hardening rates in the first few cycles were higher, with lower softening
rates in the first few cycles were higher, with lower softening rates in the following cycles under the
rates in the following cycles under the T6/T7-5 conditions than under the T7-24/T7-48 conditions,
Metals 2018, 8, 528 10 of 16

T6/T7-5
Metals conditions
2018, 8, x FORthan
PEER under
REVIEWthe T7-24/T7-48 conditions, which explains the increasing difference 10 of 15in
stress and the decreasing difference in plastic strain between the first and mid-life cycles, as shown in
which
Table 3. explains the increasing difference in stress and the decreasing difference in plastic strain
between
In the LCFthe first
tests,and
themid-life
plastic cycles, as shown inwas
strain amplitude Table 3.
considered as a physical quantity that resulted
in severalIndamaging
the LCF tests, the plastic
processes, andstrain amplitude
it influenced thewas considered
internal as a physical
microstructure, quantity
which that resulted
was closely related
in several damaging processes, and it influenced the internal microstructure, which
to the strain resistance, and eventually, the fatigue life [28]. The variation in plastic strain amplitude was closely
(∆εprelated
/2) duringto the strain resistance, and eventually, the fatigue life [28]. The variation in plastic strain
cyclic deformation at various strain amplitudes (0.4%, 0.8%, and 1.2%) under various
amplitude (  p / 2) during cyclic deformation at various strain amplitudes (0.4%, 0.8%, and 1.2%)
aging conditions is shown in Figure 8, which corresponded well with the change in stress amplitude
under
during various
cyclic aging conditions
deformation, as shownisin shown
Figurein7.Figure 8, which
As shown corresponded
in Figure wellstrain
8, the plastic with amplitude
the change also
in
stress amplitude
increased with totalduring cyclic deformation,
strain amplitude. When the as shown in Figure
total strain 7. As shown
amplitude was low in Figure
(~0.4%8,inthe plastic
Figure 8a),
strain amplitude also increased with total strain amplitude. When the total
cyclic stabilization was observed. With increasing total strain amplitude (0.8% in Figure 8b, and 1.2%strain amplitude was
low (~0.4%
in Figure in Figure
8c), plastic 8a),
strain cyclic stabilization
decreased with cycles.was observed. With increasing total strain amplitude
(0.8% in Figure 8b, and 1.2% in Figure 8c), plastic strain decreased with cycles.
It is well accepted that the cyclic hardening or softening during LCF tests is highly dependent
It is well accepted that the cyclic hardening or softening during LCF tests is highly dependent
on the ratio of σUTS to σYS , whereby cyclic hardening dominates when the ratio is higher than
on the ratio of σUTS to σYS, whereby cyclic hardening dominates when the ratio is higher than 1.4,
1.4, and cyclic softening is expected to occur when the value is lower than 1.2 [29]. As shown in
and cyclic softening is expected to occur when the value is lower than 1.2 [29]. As shown in Table 2,
Table 2, the values of the σUTS -to-σYS ratio for all four aging conditions were in the range of 1.1–1.2,
the values of the σUTS-to-σYS ratio for all four aging conditions were in the range of 1.1–1.2, and
andtherefore,
therefore,stabilization
stabilization or softening was the dominant cyclic response for all aging conditions.
or softening was the dominant cyclic response for all aging conditions. The
Thehardening
hardeningduring duringthe thefirst
firstfew
fewcycles
cycles was
was probably
probably due due to
to the
the classical
classical dislocation-dislocation
dislocation-dislocation
interactions [30,31], whereas the twisting, dissolution, and slip penetration
interactions [30,31], whereas the twisting, dissolution, and slip penetration of precipitates of precipitateswere
werethe
the
principal reasons for the softening
principal reasons for the softening [32]. [32].

Figure 8. Plastic strain amplitude vs. the number of cycles under various aging conditions tested at
Figure 8. Plastic strain amplitude vs. the number of cycles under various aging conditions tested at
various strain amplitudes of (a) 0.4%, (b) 0.8%, and (c) 1.2%.
various strain amplitudes of (a) 0.4%, (b) 0.8%, and (c) 1.2%.

3.3.3. Fatigue Life and Fatigue Fracture


3.3.3. Fatigue Life and Fatigue Fracture
Figure 9 displays the total strain amplitude ( t / 2) as a function of the number of cycles to
Figure 9 displays the total strain amplitude (∆εt /2) as a function of the number of cycles to failure
failure (Nf, i.e., the fatigue life) for the investigated 6061 alloy under various aging conditions, in
(Nf , i.e., the fatigue life) for the investigated 6061 alloy under various aging conditions, in comparison
comparison with the data reported in the literature [1]. The run-out data points are marked by
arrows pointing horizontally at or over 10 4 cycles. Overall, the investigated 6061 alloy showed a
Metals 2018, 8, 528 11 of 16

Metals
with the2018,
data8, xreported
FOR PEERin
REVIEW 11 of 15
the literature [1]. The run-out data points are marked by arrows pointing
4
horizontally at or over 10 cycles. Overall, the investigated 6061 alloy showed a trend of increasing
trend of increasing fatigue life with decreasing strain amplitude, and it also showed an improved
fatigue life with decreasing strain amplitude, and it also showed an improved fatigue life compared to
fatigue life compared to the data reported in the literature [1]. In general, the fatigue life increased
the data reported in the literature [1]. In general, the fatigue life increased with increasing T7 aging
with increasing T7 aging time. It is apparent that the fatigue life under the T7-24/T7-48 conditions
time. It is apparent that the fatigue life under the T7-24/T7-48 conditions was longer than that for
was longer than that for T6/T7-5.
T6/T7-5.

Figure9.9.Total
Figure Totalstrain amplitude
strain vs.vs.
amplitude number ofof
number cycles toto
cycles failure under
failure various
under aging
various conditions.
aging conditions.

AAsimilar
similartrend
trendwas
wasreported
reportedininthetheliterature
literature[33],
[33],ininwhich
whichthethefatigue
fatiguelife
lifeofofthe
theA356
A356alloy
alloy
was
was greatly
greatly improved
improved by prolonging
by prolonging the aging
the aging time, is
time, which which is explained
explained by the relationship
by the relationship between
between
the fatigue the
life fatigue
and yieldlifestrength.
and yield strength.
If the material If has
the amaterial
low yield has a low due
strength yield to strength
overaging, due
theto
overaging,
stress the stress
concentration concentration
generated within the generated
uneven areawithin thematerial
of the uneven isarea of relaxed
easily the material
by the is easily
plastic
relaxed by the
deformation plastic
of the part;deformation of the part;
thus, the initiation thus, the
of fatigue crack initiation of fatigue
is delayed, and itcrack
survivesis delayed, and it
for a longer
survives
period. Thefor a longer period.
relationship between Thethe
relationship
maximum betweensize of the the maximum
plastic size of the
deformation zone plastic deformation
surrounding the
zone surrounding
fatigue-crack tip andthe
the fatigue-crack
yield strengthtip andmaterial
of the the yield can strength of theasmaterial
be described follows can[33]:be described as
follows [33]:
Ktr 2
r p = 0.11( max tr) , (1)
σ𝐾
y max 2
𝑟𝑝 = 0.11( ) , (1)
𝜎𝑦
where rp is the size of the plastic zone, tr
Kmax is the transition-maximum applied stress-intensity
tr
where rp is the size of the plastic zone, K
factor, and σy is the yield strength. Based onmax is the transition-maximum
Equation (1), the low yield strength appliedcanstress-intensity
increase the
factor,
size of theand σy isplastic
local the yield strength. Based
deformation on Equation
zone, thereby (1), thethe
promoting lowcrack
yieldclosure
strength can increase
induced by thethe size
high
of the local
plasticity, plastic deformation
and subsequently, zone,
increasing thethereby promoting
crack growth the crack
resistance [33]. Inclosure induced
addition, by the
the larger high
plastic
plasticity, and
deformation zonesubsequently,
can also increaseincreasing the crack
the resistance growth
of the resistance
intracellular matrix[33].toIn addition, the
dislocations, larger
causing
plastic
these deformation
dislocations to bezone cantoalso
unable moveincrease
close tothe
theresistance
unit cell orofgrain
the intracellular
boundary to matrix
interacttowith
dislocations,
eutectic
causing therefore,
particles; these dislocations
it reducestothe
bepropagation
unable to move close
rate of to the
fatigue unit and
crack, cell further
or grainimproves
boundary thetofatigue
interact
with eutectic
properties of theparticles; therefore,the
alloy. Comparing it T6/T7-5
reduces to theT7-24/T7-48
propagationconditions
rate of fatigue crack, and further
for the investigated 6061
improves
alloy, the fatigue
the lower properties
yield strength withofprolonged
the alloy. aging
Comparing
time wasthe likely
T6/T7-5thetomain
T7-24/T7-48
reason for conditions
the longer for
the investigated
fatigue life. 6061 alloy, the lower yield strength with prolonged aging time was likely the main
reason for 10
Figure theshows
longerfracture
fatiguesurfaces
life. of the investigated alloy under the T6 and T7-48 conditions at
a total Figure 10 shows fracture
strain amplitude of 0.4%surfaces
and 1.0%, of including
the investigated alloy under
fatigue-crack the T6propagation,
initiation, and T7-48 conditions
and finalat
a total strain
fast-fracture amplitude
regions of 0.4% by
(as indicated anddashed
1.0%, yellow
includingandfatigue-crack
red lines). initiation, propagation, and final
fast-fracture regions (as indicated by dashed yellow and red lines).
Metals 2018, 8, 528 12 of 16
Metals 2018, 8, x FOR PEER REVIEW 12 of 15

Figure 10. Fatigue fracture surfaces at strain amplitudes of 0.4% and 1.0% under (a,c,e) T6 and
Figure 10. Fatigue fracture surfaces at strain amplitudes of 0.4% and 1.0% under (a,c,e) T6 and (b,d,f)
(b,d,f) T7-48 conditions.
T7-48 conditions.
It can be seen in Figure 10 that the fatigue crack initiated essentially from the specimen surface,
It can by
caused be seen in Figure 10 that
the near-surface the fatigue
defects, such ascrack initiated
porosities andessentially from the
brittle particles specimen
[34]. surface,
The greatest
caused by the near-surface defects, such as porosities and brittle particles [34]. The
differences between these two conditions (T6 and T7-48) were in the areas of the crack propagation.greatest differences
between
As shown theseintwo conditions
Figure 10b,d, the(T6 andofT7-48)
area were
the crack in the areas
propagation of the
under thecrack
T7-48propagation.
conditions was Asmuch
shown
larger than
in Figure that
10b,d, theunder thethe
area of T6crack
conditions (Figure under
propagation 10a,c) the
at both
T7-48low (0.4%) and
conditions washigh (1.0%)
much strain
larger than
amplitude, which was corroborated by the larger plastic deformation (Figure 6) and
that under the T6 conditions (Figure 10a,c) at both low (0.4%) and high (1.0%) strain amplitude, which longer fatigue
waslife of the T7-48 by
corroborated sample (Figure
the larger 9) compared
plastic to the(Figure
deformation T6 sample.
6) andFatigue
longerstriations on the
fatigue life fatigue
of the T7-48
fracture
sample surface
(Figure under both
9) compared to conditions wereFatigue
the T6 sample. also observed, especially
striations at lower
on the fatigue stain surface
fracture amplitudes,
under
such
both as 0.4%, were
conditions as shown
also in Figure 10e,f.
observed, The presence
especially at lowerofstain
fatigue-striation
amplitudes, marks
such as is 0.4%,
characteristic
as shown of in
the crack-propagation regime in ductile materials [17], such as the investigated 6061
Figure 10e,f. The presence of fatigue-striation marks is characteristic of the crack-propagation regime in aluminum
alloys. It was observed that the spacing of fatigue striations in the T6 sample was somewhat larger
ductile materials [17], such as the investigated 6061 aluminum alloys. It was observed that the spacing
than that in the T7-48 sample, indicating a faster crack-propagation rate. This explains the shorter
of fatigue striations in the T6 sample was somewhat larger than that in the T7-48 sample, indicating
fatigue life in the T6 sample, combined with a smaller propagation region.
a faster crack-propagation rate. This explains the shorter fatigue life in the T6 sample, combined with
a smaller propagation
3.3.4. Assessment region. Parameters
of Fatigue
Based on the
3.3.4. Assessment ofLCF results
Fatigue (Figures 6 and 8), the fatigue parameters were calculated to assess the
Parameters
fatigue life of the investigated 6061 aluminum alloy under various conditions. The total strain
Based on the LCF results (Figures 6 and 8), the fatigue parameters were calculated to assess
amplitude could be expressed as two parts: the plastic and elastic strain amplitudes, according to the
the fatigue life of the investigated 6061 aluminum alloy under various conditions. The total strain
Coffin-Manson and Basquin relationships [35], i.e.,
Metals 2018, 8, 528 13 of 16

amplitude could be expressed as two parts: the plastic and elastic strain amplitudes, according to the
Coffin-Manson and Basquin relationships [35], i.e.,
 b
0 2N
∆ε t ∆ε p ∆ε e   c σ f f
= + = ε f 0 2N f + , (2)
2 2 2 E
where E is the Young’s modulus (for the presented alloy, the average value obtained during fatigue
testing was ~68 GPa), Nf is the fatigue life or the number of cycles to failure, σ’f is the fatigue strength
coefficient, b is the fatigue strength exponent, ε’f is the fatigue ductility coefficient, and c is the fatigue
ductility exponent. In addition, cyclic deformation behavior is normally considered to be related to the
portion of plastic strain amplitude, and is independent of the elastic strain amplitude, which can be
expressed by the following equation [36]:
n0
∆σ 0 ∆ε p

=K (3)
2 2

∆ε
where ∆σ p
2 is the mid-life stress amplitude, 2 is the mid-life plastic strain amplitude, n’ is the cyclic
strain-hardening exponent, and K’ is the cyclic strength coefficient. The estimated LCF parameters
based on Equations (2) and (3), and on the LCF results are presented in Table 4, and are compared
with the data reported in the literature for extruded 6061-T6 alloys [1,9]. To ensure that the cyclic
stabilization, often called cyclic saturation, already occurred, the stress and strain values of the mid-life
cycle were used for the calculations.
As shown in Table 4, it can be seen that the estimated fatigue parameters were well within the
range of other extruded 6061 alloys reported in the literature [1,9]. The T6 alloy had the highest σ’f and
b values, but the lowest ε’f and c values of all the examined conditions. This indicates that alloys with
higher σ’f and b values have higher strength, whereas those with higher ε’f and c values have higher
ductility [2], which also coincides with the explanation for the longer plastic deformation for the T7
aged alloys (Figure 5). It was observed that the value of the cyclic strain-hardening exponent, n’, of the
T6 alloy was lower than that of the alloys after T7 aging, such as T7-24 and T7-48. It was also reported
in the work of Nandy et al. [17] that the cyclic strain-hardening exponent increased from under-aged
to peak-aged conditions, and further increased to over-aged conditions. In addition, the values of n’ in
Table 4 during cyclic deformation were higher than its monotonic strain-hardening values n in Table 2,
which directly reflects a higher cyclic stress than monotonic tensile stress at the same strain for all
conditions tested. Moreover, both the cyclic strength coefficient, K’, and fatigue strength coefficient, σ’f ,
decreased with increasing T7 aging time, leading to the lowest stress amplitude (Figure 5 and Table 3)
in the T7-48 alloy. It is worth mentioning that the fatigue ductility coefficient, ε’f , was found to increase
with increasing T7 aging time, indicating a higher plastic strain with longer aging time, and therefore,
the longer fatigue life of the alloy under the T7-24 and T7-48 conditions compared to that under the
T6 conditions.

Table 4. Low-cycle fatigue (LCF) parameters estimated for 6061 aluminum alloys under
various conditions.

Low-Cycle Fatigue Parameters T6 T7-5 T7-24 T7-48 T6 [1] T6 [8]


Cyclic strain-hardening exponent, n’ 0.14 0.12 0.15 0.18 0.078 0.24
Cyclic strength coefficient, K’ (MPa) 636 621 552 478 268 372
Fatigue strength coefficient, σ’f (MPa) 872 613 602 458 705 593
Fatigue strength exponent, b −0.087 −0.110 −0.110 −0.214 −0.11 −0.093
Fatigue ductility coefficient, ε’f 0.70 0.90 1.80 4.49 2.40 5.39
Fatigue ductility exponent, c −1.10 −0.87 −0.80 −0.79 −0.98 −1.10
Metals 2018, 8, 528 14 of 16

4. Conclusions
Strain-controlled low-cycle fatigue tests were conducted on an extruded AA6061 aluminum alloy
at varying strain amplitudes to determine the effect of overaging on the cyclic deformation behavior.
The following conclusions could be drawn:

(1) During the T7 aging treatment at 200 ◦ C, no remarkable changes in the microstructure (Al-Fe-Si
and Mg2 Si intermetallics, and α-Al grain size) could be observed with increasing aging time.
However, the size of precipitates increased from 60 nm under the T6 conditions to 220 nm after
aging 48 h at 200 ◦ C, leading to a decrease in the monotonic tensile strength.
(2) The hysteresis loops of the first cycle after a long T7 aging time were moderately inflected
compared to those under the T6 conditions, whereas nearly symmetrical loops were present
during the mid-life cycle under all conditions tested. The peak stresses decreased and the plastic
strain increased with increasing T7 aging time. Nearly ideal Masing behavior was exhibited
under the T6 conditions, whereas it was lost under the overaging conditions.
(3) Similar cyclic stress responses were found under all tested conditions: cyclic stabilization was
present at low strain amplitudes (0.2–0.4%), whereas cyclic softening with initial hardening
during the first few cycles occurred at middle-to-high strain amplitudes (0.6–1.2%). The softening
rate increased gradually with increasing T7 aging time.
(4) Compared to the T6 conditions, the fatigue life increased with increasing T7 aging time.
The estimated fatigue parameters indicated that the fatigue ductility coefficient increased and the
fatigue strength coefficient decreased with increasing T7 aging time.

Author Contributions: K.L., F.A.M. and X.G.C. conceived and designed the experiments; F.A.M. and K.L.
performed the experiments; F.A.M. analyzed the data; F.A.M. and K.L. wrote the paper; and X.G.C. modified
the paper.
Funding: This research received no external funding.
Acknowledgments: The authors would like to thank the Natural Sciences and Engineering Research Council
of Canada (NSERC) and Rio Tinto Aluminum, through the NSERC Industrial Research Chair in Metallurgy of
Aluminum Transformation at the University of Québec at Chicoutimi (UQAC) for providing financial support.
The authors would also like to thank Q. Li (Ryerson University) and D. Racine and P.-L. Privé (UQAC) for their
assistance in the experiments.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Brammer, A.T.; Jordo, J.B.; Allison, P.G.; Barkey, M.E. Strain-controlled low-cycle fatigue properties of
extruded 6061-T6 aluminum alloy. J. Mater. Eng. Perform. 2013, 22, 1348–1350. [CrossRef]
2. Abood, A.N.; Saleh, A.H.; Abdullah, Z.W. Effect of heat treatment on strain life of aluminum alloy AA6061.
J. Mater. Sci. Res. 2013, 2, 51–59.
3. Yassar, R.S.; Field, D.P.; Weiland, H. Transmission electron microscopy and differential scanning calorimetry
studies on the precipitation sequence in an Al-Mg-Si alloy: AA6022. J. Mater. Res. 2011, 20, 2705–2711.
[CrossRef]
4. Polat, A.; Avsar, M.; Ozturk, F. Effects of the artificial-aging temperature and time on the mechanical
properties and springback behavior of AA6061. Mater. Technol. 2015, 49, 487–493. [CrossRef]
5. Buha, J.; Lumley, R.N.; Crosky, A.G. Microstructural development and mechanical properties of interrupted
aged Al-Mg-Si-Cu alloy. Metall. Mater. Trans. A 2006, 37, 3119–3130. [CrossRef]
6. Haji, Z.N. Low cycle fatigue behavior of aluminum alloys AA2024-T6 and AA7020-T6. J. Eng. Sci. 2010, 22–23,
127–137.
7. Sharma, V.; Rao, G.; Sharma, S.; George, K. Low cycle fatigue behavior of AA2219-T87 at room temperature.
Mater. Perform. Charact. 2014, 3, 103–126. [CrossRef]
8. Hidayetoglu, T.K.; Pica, P.N.; Haworth, W.L. Aging dependence of the bauschinger effect in aluminum alloy
2024. Mater. Sci. Eng. 1985, 73, 65–76. [CrossRef]
Metals 2018, 8, 528 15 of 16

9. Wong, W.A.; Bucci, R.J.; Stentz, R.H.; Conway, J.B. Tensile and Strain-Controlled Fatigue Data for Certain
Aluminum Alloys for Application in the Transportation Industry; SAE International: Warrendale, PA, USA, 1987.
10. Srivatsan, T.S.; Kolar, D.; Magnusen, P. Influence of temperature on cyclic stress response, strain resistance,
and fracture behavior of aluminum alloy 2524. Mater. Sci. Eng. A 2001, 314, 118–130. [CrossRef]
11. Borrego, L.P.; Abreu, L.M.; Costa, J.M.; Ferreira, J.M. Analysis of low cycle fatigue in almgsi aluminium
alloys. Eng. Fail. Anal. 2004, 11, 715–725. [CrossRef]
12. Lapovok, R.; Loader, C.; Torre, F.H.D.; Semiatin, S.L. Microstructure evolution and fatigue behavior of
2124 aluminum processed by ECAE with back pressure. Mater. Sci. Eng. A 2006, 425, 36–46. [CrossRef]
13. Jogi, B.F.; Brahmankar, P.K.; Nanda, V.S.; Prasad, R.C. Some studies on fatigue crack growth rate of aluminum
alloy 6061. J. Mater. Process. Technol. 2008, 201, 380–384. [CrossRef]
14. Tsuyoshi, T.; Sasaki, K. Low cycle thermal fatigue of aluminum alloy cylinder head in consideration of
changing metrology microstructure. Procedia Eng. 2010, 2, 767–776. [CrossRef]
15. Zhu, M.; Jian, Z.; Yang, G.; Zhou, Y. Effects of T6 heat treatment on the microstructure, tensile properties,
and fracture behavior of the modified A356 alloys. Mater. Des. 2012, 36, 243–249. [CrossRef]
16. Fan, K.L.; He, G.Q.; Liu, X.S.; Liu, B.; She, M.; Yuan, Y.L.; Yang, Y.; Lu, Q. Tensile and fatigue properties of
gravity casting aluminum alloys for engine cylinder heads. Mater. Sci. Eng. A 2013, 586, 78–85. [CrossRef]
17. Nandy, S.; Sekhar, A.P.; Kar, T.; Ray, K.K.; Das, D. Influence of ageing on the low cycle fatigue behaviour of
an Al–Mg–Si alloy. Philos. Mag. 2017, 97, 1978–2003. [CrossRef]
18. Su, L.; Lu, C.; Deng, G.; Tieu, A.K.; Sun, X. Microstructure and mechanical properties of 1050/6061 laminated
composite processed by accumulative roll bonding. Rev. Adv. Mater. Sci. 2013, 33, 33–37.
19. Osten, J.; Milkereit, B.; Schick, C.; Kessler, O. Dissolution and precipitation behaviour during continuous
heating of Al–Mg–Si alloys in a wide range of heating rates. Materials 2015, 8, 2830–2848. [CrossRef]
20. Edwards, G.A.; Stiller, K.; Dunlop, G.L.; Couper, M.J. The precipitation sequence in Al–Mg–Si alloys.
Acta Mater. 1998, 46, 3893–3904. [CrossRef]
21. Ozturk, F.; Sisman, A.; Toros, S.; Kilic, S.; Picu, R.C. Influence of aging treatment on mechanical properties of
6061 aluminum alloy. Mater. Des. 2010, 31, 972–975. [CrossRef]
22. Liu, K.; Chen, X.G. Development of Al-Mn-Mg 3004 alloy for applications at elevated temperature via
dispersoid strengthening. Mater. Des. 2015, 84, 340–350. [CrossRef]
23. Ebrahimi, R.; Pardis, N. Determination of strain-hardening exponent using double compression test.
Mater. Sci. Eng. A 2009, 518, 56–60. [CrossRef]
24. Abu-Haiba, M.S.; Fatemi, A.; Zoroufi, M. Creep deformation and monotonic stress-strain behavior of haynes
alloy 556 at elevated temperatures. J. Mater. Sci. 2002, 37, 2899–2907. [CrossRef]
25. Suresh, S. Fatigue of Materials; Cambridge University Press: Cambridge, UK, 2006.
26. Wang, Z.; Laird, C. Relationship between loading process and masing behavior in cyclic deformation. Mater.
Sci. Eng. A 1988, 101, L1–L5. [CrossRef]
27. Christ, H.J.; Mughrabi, H. Cyclic stress-strain response and microstructure under variable amplitude loading.
Fatigue Fract. Eng. Mater. Struct. 1996, 19, 335–348. [CrossRef]
28. Mirza, F.A.; Chen, D.L.; Li, D.J.; Zeng, X.Q. Low cycle fatigue of an extruded Mg-3Nd-0.2Zn-0.5Zr magnesium
alloy. Mater. Des. 2014, 64, 63–73. [CrossRef]
29. Xue, L. A unified expression for low cycle fatigue and extremely low cycle fatigue and its implication for
monotonic loading. Int. J. Fatigue 2008, 30, 1691–1698. [CrossRef]
30. Lam, P.C.; Srivatsan, T.S.; Hotton, B.; Al-Hajri, M. Cyclic stress response characteristics of an aluminum-
magnesium-silicon alloy. Mater. Lett. 2000, 45, 186–190. [CrossRef]
31. Calabrese, C.; Laird, C. Cyclic stress—Strain response of two-phase alloys Part I. Microstructures containing
particles penetrable by dislocations. Mater. Sci. Eng. 1974, 13, 141–157. [CrossRef]
32. Bhat, S.P.; Laird, C. High temperature cyclic deformation of precipitation hardened alloy—I. Partially
coherent precipitates. Acta Metall. 1979, 27, 1861–1871. [CrossRef]
33. Song, M.S.; Ran, M.W.; Kong, Y.Y.; Yan, D.Y. Low cycle fatigue behavior of cast A356 aluminum alloys.
Chin. J. Nonferr. Met. 2011, 21, 538–545.
34. Tian, D.D.; Liu, X.S.; He, G.Q.; Shen, Y.; Lv, S.Q.; Wang, Q.G. Low cycle fatigue behavior of casting A319
alloy under two different aging conditions. Mater. Sci. Eng. A 2016, 654, 60–68. [CrossRef]
Metals 2018, 8, 528 16 of 16

35. Mirza, F.A.; Chen, D.L. Fatigue of magnesium alloys. In Aerospace Materials Handbook; Zhang, S., Zhao, D.L.,
Eds.; CRC Press: Boca Raton, FL, USA; Taylor& Francis: New York, NY, USA, 2013; pp. 647–698.
36. Begum, S.; Chen, D.L.; Xu, S.; Luo, A.A. Strain-controlled low-cycle fatigue properties of a newly developed
extruded magnesium alloy. Metall. Mater. Trans. A 2008, 39, 3014–3026. [CrossRef]

© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

Das könnte Ihnen auch gefallen