Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

International Journal of Fatigue 109 (2018) 1–9

Contents lists available at ScienceDirect

International Journal of Fatigue


journal homepage: www.elsevier.com/locate/ijfatigue

Grain size dependence of fatigue properties of friction stir processed T


ultrafine-grained Al-5024 alloy

Shivakant Shukla, Mageshwari Komarasamy, Rajiv S. Mishra
Center for Friction Stir Processing, Advanced Materials and Manufacturing Processes Institute, Department of Materials Science and Engineering, University of North Texas,
Denton, TX 76207, USA

A R T I C L E I N F O A B S T R A C T

Keywords: The main objective of the present study was to examine the effect of grain size on the fatigue properties of 5024
Ultrafine grain Al alloy. Two ultrafine-grained (UFG) and one fine-grain microstructures were produced via friction stir pro-
Fatigue properties cessing (FSP). Fatigue properties were investigated using a fully reversible bending fatigue testing machine.
Friction stir processing Microstructural evolution during fatigue was analyzed by carrying out interrupted fatigue testing with sub-
Basquin equation parameters
sequent orientation imaging microscopy. Emphasis was given to grain size on the evolution of deformed
structure, intergranular dislocation density, and overall S-N curve. A comparison with literature demonstrates
the importance of the processing route to obtain the UFG microstructure. This difference is highlighted through
an analysis of Basquin equation parameters. The results indicate that FSP can be utilized as an effective route to
obtain stable UFG microstructure. A fractographic study was done to understand the effect of microstructure on
fatigue crack initiation and propagation rates. Crack growth rates in stage-II has been compared for the three
different grain sizes.

1. Introduction aforementioned SPD processes have high fractions of non-equilibrium


grain boundaries and increased dislocation densities when compared
Improvement in tensile properties, which has long been known to with FSP [11,12]; the result is microstructural instability, i.e., the for-
enhance fatigue properties of material, has led to the development of a mation of shear bands (SBs) and grain/cell coarsening. While tensile
number of strengthening mechanisms that rely on a combination of properties of UFG materials do increase by grain refinement using other
solid solution, precipitates, and grain refinement [1,2]. Considerable SPD methods, the above-mentioned microstructural instabilities bring
effort has been expended on grain refinement to improve material new challenges to fatigue properties. These microstructural instabilities
strength. Severe plastic deformation (SPD) methods such as equal lead to cyclic softening, which detrimentally affects fatigue properties.
channel angular pressing [3], high pressure torsion [3], accumulating Many studies confirm that the need for annealing heat treatment to
roll bonding [4], and multiaxial forging [5] are effective techniques to improve fatigue properties in SPD-processed materials [13,14]. While
obtain bulk ultra-fine grained (UFG) microstructures (grain size < tensile properties are generally affected by the overall characteristics of
1 µm). A different approach is the ball-milling of powders, preferably at the microstructure, fatigue properties are influenced by local micro-
cryogenic temperatures, to suppress recovery and subsequent re- structural imperfections such as microstructural instability and grain/
crystallization processes leading to UFG microstructure [6,7]. This ap- cell coarsening. Thus, a direct relation between tensile and fatigue
proach is generally not suitable for bulk UFG materials; hence SPD properties of UFG materials is difficult to ascertain. In this regard, FSP
processes remain by far the most suitable method [8]. Another such can be a viable technique to overcome these challenges, as dynamic
SPD method is friction stir processing (FSP), a derivative of friction stir recrystallization involved in FSP engenders the formation of fine grains
welding (FSW) developed by Mishra et al. [9], where the material is with a high fraction of high angle grain boundaries (HAGB) and low
severely deformed by a rotating pin with shoulder with a rise in dislocation density. These characteristics are beneficial for fatigue
workpiece temperature high enough for recrystallization to occur. A properties [15]. One long-established fact about cyclic loading is that
variation in dislocation density and grain boundary character dis- deformation in coarse grain (CG) materials occurs through nucleation
tribution for each kind of SPD method affects both tensile and cyclic and multiplication of dislocations within the grain. Dislocation density
loading properties of the material [3,10]. Microstructures produced via continues to increase, and deformation leads to the rearrangement and


Corresponding author.
E-mail address: rajiv.mishra@unt.edu (R.S. Mishra).

https://doi.org/10.1016/j.ijfatigue.2017.12.007
Received 19 September 2017; Received in revised form 17 November 2017; Accepted 9 December 2017
Available online 11 December 2017
0142-1123/ © 2017 Elsevier Ltd. All rights reserved.
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

formation of three-dimensional boundaries that, on further application regime, a step size of 80 nm was used. Interrupted fatigue testing on
of stress, change into a two-dimensional (2-D) shape called sub-grain UFG410 with subsequent OIM analysis with the above-mentioned
boundaries [16]. Size and shapes of these sub-grain boundaries or cells parameter was performed to understand microstructural evolution
generally depend on the material and applied strain. For pure alu- during fatigue testing. Fractography was done using FEI Quanta ESEM.
minum, this typical subgrain structure is ∼400 nm [17]. Given that For fractographic studies, samples were preserved immediately, and
grain sizes in UFG Al alloys are comparable to or smaller than the ty- minimal cleaning was done.
pical dislocation substructure size, a question arises as to how the given
length scale affects the formation of sub-grain cells and the interaction 3. Results and discussion
of dislocations with boundaries becomes important.
Because of their lightweight, aluminum alloys have been used ex- 3.1. Initial microstructure characterization
tensively for the structural applications in the aerospace industries. The
addition of magnesium increases strength via solid solution strength- Initial EBSD analysis of all the three samples was performed to
ening without compromising the excellent strength-to-weight ratio. obtain grain size, grain size distribution, and grain boundary mis-
Further micro-alloying addition of Sc (< 0.4%) improves strength sig- orientation angle distribution. The OIM micrographs of 5024 Al
nificantly due to the formation of coherent nanoscale Al3Sc precipitates UFG410, UFG210, and FG alloys with their corresponding GSD are shown
[18,19]. Al3Sc precipitates also hinder the mobility of grain boundaries in Fig. 1. Inverse pole figure (IPF) in Fig. 1(g) shows crystallographic
through Zener pinning and thus serves as a grain-refining agent during orientation associated with different colors. Table 2 average value of
high-temperature processing and application. the grain size of UFG410, UFG210 and FG materials.
The aim of the present study was to understand how grain size In UFG410 material, 98% of grains were < 1 µm, and 25% of grains
variation in UFG regime affects fatigue properties. Thus, two different were below 500 nm; while in UFG210 material, 95% grains were < 1
UFG microstructures were obtained with a very high fraction of HAGB. µm, and 72% were < 500 nm. Of note is that step size was 80 nm. Even
The difference between the cyclic loading behavior of UFG and fine- though smaller step size would yield more information about the GSD,
grained 5024 Al alloy was investigated in terms of S-N curve and it should not have any major effect on average grain size information.
fractography. On the other hand, average grain size for FG material was ∼4.61 µm,
with 93% grains below 10 µm and only 2% grains below 1 µm. Al3Sc
precipitates in 5024 Al provide excellent thermal stability to grain
2. Experimental growth [22]. Addition of Sc and Zr also suppresses abnormal grain
growth up to a temperature of 823 K [22].
Al-Mg-Sc alloy of 7 mm thickness was obtained from Aleris with a
nominal composition of Al-4.0 wt% Mg-0.3 wt% Sc. FSP was done on 3.2. Tensile properties
the plate to obtain UFG microstructure. Details of the tool used in this
processing were reported in [20,22]. Two different grain sizes obtained Typical stress – strain curve for all the three microstructural con-
by varying the FSP parameters will be referred to as UFG410 and UFG210 ditions are shown in Fig. 2. Yield strength (YS) and ultimate tensile
alloys as detailed in Table 1. stress (UTS) for UFG410 were 320 ± 3 MPa and 350 ± 5 MPa, re-
During UFG210 alloy processing, a copper backing plate was used to spectively with uniform elongation of ∼12% and total elongation of
accelerate heat flow so that a faster cooling rate could be obtained. 20%. For UFG210 5024 Al alloy, YS and UTS values were 410 ± 5 MPa
Apart from the two UFG 5024 Al materials, a fine grain microstructure and 415 ± 5 MPa, respectively, with ∼3% uniform elongation and 6%
(average grain size > 1 µm) was produced by annealing the UFG410 failure elongation; while FG 5024 Al showed UTS of 290 ± 5 MPa and
alloy at 873 K for 1.5 h and hereinafter referred to as FG. FG alloy was YS of 175 ± 3 MPa, with uniform elongation of ∼14% and 16% total
obtained to compare the fatigue properties of UFG materials with fine- elongation. As expected, the decrease in grain size has a significant
grain one. Tensile tests were carried out for all three conditions using a effect on tensile properties. YS of UFG alloy increased by almost
computer-controlled mini-tensile testing machine at an initial strain 100 MPa when average grain size decreased from ∼ 410 nm
rate of 10−3 s−1. Gage length of the samples was ∼5 mm with width to ∼ 210 nm.
∼1.2 mm, and thickness ∼1.1 mm. Mechanical polishing of all tensile The significant impact of grain size on uniform elongation and work
samples was carried out to 1 µm surface finish using diamond suspen- hardening indicates that an increase in grain size improves work
sion on Kempad cloth. Three samples were tested for statistical vali- hardening of 5024 Al alloy. This is related to the dislocation storage
dation. Bending fatigue testing was performed using a custom-made when grain size reaches UFG domain [15]. Also, worth noting is the
tabletop mini-fatigue testing machine; complete details are in [23]. All variation in serration behavior observed in these alloys as a function of
samples were tested at 20 Hz frequency and stress ratio (R) of −1. Since grain size. As compared to FG alloy, serration amplitude (Δσ) in UFG410
fatigue properties are affected by surface condition, sample preparation alloy was higher and initiated at the beginning of plastic deformation.
was done with extreme precaution. Samples were polished down to a Komarasamy and Mishra reported that the serreation amplitude and
surface finish of 1 µm by diamond suspension followed by 0.05 µm, frequency increased with decrease in grain size [20]. Such observations
then 0.02 µm surface finish using colloidal silica suspension to elim- have been attributed to decreasing grain size and resultant change in
inate surface flaws. Grain size distribution (GSD) and misorientation the interaction between the dislocations and various microstructural
angle distribution information was obtained by orientation image mi- elements (solutes and grain boundaries). The above-mentioned values
croscopy (OIM) on FEI Nova NanoSEM 230, with a 15 kV accelerating of YS for all three microstructures followed a Hall-Petch relationship
voltage. The EBSD analysis was done using SL Digiview III Electron [21], and the calculated grain size of 5024 Al UFG210 alloy by this
Backscatter Diffraction (EBSD) detector. Since the grains fall in the UFG method is 198 nm, which was reasonably close to the value obtained via
OIM analysis (∼210 nm).
Table 1
Details of FSP parameters for processing AA2054 alloy.
3.3. Fatigue behavior
Microstructure Rotational rate Traverse speed Tool tilt Cooling
(RPM) (IPM) (Degrees) To understand the overall fatigue behavior of 5024 Al UFG410,
UFG210, and FG alloys, S-N curves were generated for all three micro-
UFG410 325 8 -2.5 No cooling
UFG210 275 8 -2.5 Cu backing structural conditions at R of −1. Fig. 3(a) presents S-N data for all three
conditions. At a particular stress amplitude (e.g. at 220 MPa), UFG210

2
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 1. OIM micrographs of 5024 Al (a) UFG410, (b) UFG210, and (c) FG material and corresponding grain size distribution (d), (e), and (f) with (g) inverse pole figure.

alloy with the smallest grain size performed better as compared to


Table 2 UFG410 alloy, thereby confirming that finer grain size in the UFG re-
Average grain size values for UFG410, UFG210, and FG condition. gime does lead to longer fatigue life.
Although S-N curve for FG alloy was situated below both UFG410
Material UFG410 UFG210 FG and UFG210 material, its proximity with UFG410 S-N curve suggests that
Grain size 0.410 ± 0.280 µm 0.210 ± 0.110 µm 4.61 ± 3.45 µm fatigue properties of FG material are very close to UFG410 condition.
The fatigue curves of the FG and the UFG410 microstructures are very
close to each other, but the scatter in the FG data is comparatively

Fig. 2. (a) Engineering stress-strain curves and (b) true stress-strain curves for 5024 Al UFG410, UFG210, and FG alloys.

3
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 3. (a) S-N curves for 5024 Al UFG410, UFG210, and FG alloys and (b) normalized S-N curves for all three conditions.

large. Also, because of the lower yield strength of the FG micro- variation, and local misorientation change. Fig. 4(c) and (d) shows a
structure, the stress range of testing is quite narrow. Since all the three variation in grain size and misorientation angle before fatigue testing
microstructures have different UTS values comparison of their fatigue and after 15,000 cycles. The average grain size changed from
behavior on UTS normalized basis can provide additional insight 410 ± 280 nm to around 430 ± 350 nm after 15,000 cycles. The
(Fig. 3(b)). Comparison of UFG410 with UFG210 reveals that at higher above observation confirms that only minimal grain variation, if any
normalized stress (> 0.5 UTS), UFG410 performs better than UFG210 occurred during fatigue testing. The local grain coarsening during fa-
condition; while at lower stresses (around 0.4 UTS), UFG210 performs tigue, which is a characteristic feature of SPD processed material [16],
almost equal to UFG410 condition. Note also that 5024 Al FG material in was not observed. Fig. 4(d) presents the variation in misorientation
normalized stress plot performs better than both UFG410 and UFG210 before fatigue testing and after 15,000 cycles; clearly, the fraction of
microstructures. According to Landgraf [24] and Morrow [25], fatigue HAGB was also quite high indicating that the microstructure was fully
properties can be categorized into low cycle fatigue (LCF) and high recrystallized and stable. Although there was a minor variation in
cycle fatigue (HCF), and are determined by both ductility and strength. misorientation angle after 15,000 cycles, the fraction of HAGB re-
LCF is governed by fatigue ductility, while HCF is controlled by fatigue mained almost the same even after 15,000 cycles.
strength [16,24,25]. Since UFG210 alloy has higher tensile strength and In the present study, kernel average misorientation (KAM) analysis
inferior ductility than UFG410 alloys, UFG210 alloy can be expected to has been used to qualitatively compare the microstructure before and
perform better in a high cyclic regime; while UFG410 alloy will be better after fatigue (15,000 cycles). This approach is similar to the reports by
suited for lower cycle regime of HCF. A recent study by Liu et al. on Cu- Ulacia et al. [26], Yi et al. [27], and Calcagnotto et al. [28], where KAM
15 at% Al alloy also shows similar observations for high cycle region values have been related to dislocation density. The objective is to
[1]. understand change in local misorientation as KAM analysis can be a
useful tool in quantifying the average misorientation around a point
3.4. Interrupted fatigue testing and subsequent OIM analysis with respect to a specified set of nearest or nearest and next-nearest
neighbor points (here 2nd nearest neighbor to reduce the noise). Data
Grain coarsening has been a prime microstructural phenomenon in above the predefined threshold value (2° for present study) are ex-
UFG alloys during cyclic loading. Grain coarsening is generally believed cluded from the calculation, since misorientations in the range of 2–15°
to occur by dynamic recrystallization, and may be the indication of are assigned to subgrain boundaries. Fig. 5(a) and (b) shows the change
fatigue damage of the microstructure [16]. To study microstructural in KAM before and after 15,000 cycles, respectively. All the other
changes during fatigue testing, 5024 Al UFG410 specimen was subjected parameters such as step size, frame, and nearest neighbors in KAM
to interrupted fatigue test and subsequent OIM analysis. analysis were kept same. To ensure a more accurate comparison, the
color assigned to each orientation of grains in both micrographs was
1. OIM was done at a particular location in the gauge section before also kept same.
fatigue. For certainty of the location, multiple images were taken at The total fraction above 0.8° misorientation (i.e., from 0.8° to 2°
various magnifications. misorientation) increased from 0.40 to 0.46; and the fraction above 1.2°
2. The specimen was subjected to a stress level of 265 MPa and was almost doubled after 15,000 cycles. This certainly points towards in-
interrupted at 15,000 cycles, the number of cycles required for crack tragranular dislocation storage, and is an indication of excellent cyclic
initiation. This stress and number of cycles were selected based on stability. Note that the grains in UFG410 are higher than the critical
the results from previous testing at the same stress amplitude. grain size for the intragranular storage [17].
3. After 15,000 cycles, OIM measurements were done at the same lo- To better understand the reason for this fatigue behavior, the S-N
cation using the same scanning parameters; sample orientation was curve in Fig. 3 was fitted with Basquin equation, Δσ = σf′·(2Nf)b, where
also kept the same. σf′ is the fatigue strength coefficient, which is the intercept of the S-N
4. The sample was then loaded again at the same stress amplitude and curve, and b is the slope of S-N curve. To enable a better analysis and
tested until specimen failure, which was around 25,000 cycles. more objective comparison, three more datum points from the work of
Vinogradov et al. [29] were considered. Fig. 6 shows the relationship
Fig. 4(a) and (b) shows OIM micrographs taken before fatigue and among these parameters with respect to grain size. Fig. 6(a) and (b)
after 15,000 cycles, respectively. The same grain was circled in both shows fatigue strength coefficient (σf′) and fatigue strength (σ−1) as a
images for comparison. From the above test data analyses was per- function of grain size. It can be seen that for the present study, both
formed regarding grain size variation, grain boundary character fatigue strength coefficient (σf′) and fatigue strength (σ−1) decrease

4
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 4. (a) and (b) OIM micrographs of same location before fatigue and after 15,000 cycles, (c) grain size variation and (d) misorientation angle variation.

continuously with an increase in grain size. Since tensile strength with an increase in grain size. No such correlation can be noted in the
generally governs fatigue strength coefficient (σf′) and decreasing grain work of Vinogradov et al. [29]. Fatigue strength exponent (b) in
size increases tensile strength, a decrease in both σf′ and σ−1 is expected Fig. 6(c), which indicates the damage mechanism [30], follows almost

Fig. 5. KAM analysis for dislocation storage on


UFG410 sample tested at a stress level of 265 MPa,
(a) before fatigue and (b) after 15,000 cycles.

5
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 6. Illustration of fatigue parameters as a function of grain size. (a) Fatigue strength exponent and grain size, (b) fatigue strength coefficient and grain size, (c) fatigue strength and
grain size, (d) fatigue ratio and grain size. Al–4.5 Mg–0.2Sc–0.2Zr, Al–3 Mg–0.2Sc–0.18Zr, and Al–1.5 Mg–0.2Sc–0.2Zr data points have been taken from Vinogradov et al. [29].

the same trend as does the (σ−1/σUTS) fatigue ratio (Fig. 6(d)) with cracks or crack nucleation points. The macroscopic view of the frac-
respect to the grain size. For the present study, it can be seen that with tured surface has three distinct regions. Region 1 (Fig. 7(a)) was gen-
an increase in grain size, fatigue strength exponent (b) increases erally smooth at low magnification and included crack initiation points
slightly, while the datum points from Vinogradov el al. [29] don’t as well as stage I of crack growth [31]. Features such as river mark or
follow this trend. We propose that the reason behind such observations beach marks as well as the orientation of fatigue striation are important
is the difference between the inherent mechanisms of grain refinement in identifying the crack initiation [33]. Fig. 7(b) is a higher magnifi-
during FSP. In the present work, the UFG microstructure was obtained cation image of the same region consisting of ridges and plateaus. In
via FSP while ECAP was used to produce UFG microstructure in the this region, the crack grew mainly along crystallographic directions.
work of Vinogradov et al. [29]. UFG alloys produced by FSP technique Region 2 is stage II of crack propagation and often resulted in crack-
contain a high fraction of equilibrium high angle boundaries because of arrest marks, also known as fatigue striations, on the fracture surface.
dynamic recrystallization during FSP. Further, due to the high thermal Such crack-arrest marks are generally the indication of crack front
energy involved, the intra-granular dislocation density is lower [11,12]. position during this stage of propagation.
The UFG microstructure produced via ECAP and other SPD techniques Region 3 was the overload or final failure zone that occurred by
contain high dislocation density within the subgrains [12] as the grain void nucleation, coalescence and growth; and consisted of a combi-
refinement mechanism involves fragmentation of grains by creation of nation of coarse and fine dimples. Fig. 7(c) shows stage II and stage
dislocation substructure and subsequent conversion to high angle grain III of fatigue failure; while 7(d) shows the typical striation marks
boundaries. Also, worth noting is the high value of the fatigue ratio (b) indicating stage II of crack growth regime. The average distance
for FG condition microstructure. This can be correlated to excellent between striation marks was ∼1.8 μm to 2 μm. Williams et al. [32]
damage tolerance ability of the fine grain material leading to a high conducted experiments on AA7075-T6 alloy and concluded that in
fatigue strength ratio value. stage II crack growth regime, the association between striation spa-
cing and crack growth rate in that striation spacing is generally
3.5. Fractography equivalent to the crack growth rate (da/dN). Apart from this, Pelloux
et al. [33], with their experiment on AA2124–T351, demonstrated
Fractography was done to observe the features of fatigue deforma- that each cycle imposes one striation and width typically in corre-
tion. Samples from all three conditions were examined under SEM to spondence with cycle amplitude. Wanhill and Hattenberg [34] and
determine the difference in various modes of crack propagation and Wanhill and Molent [35] have used these markings to determine the
their rates. To conduct a more comprehensive analysis, stress value crack growth rate in aircraft component. A recent review article by
with respect to UTS (i.e., normalized stress) was kept close to ∼0.7. Lynch [36] mentions that fatigue crack growth for intermediate ΔK
Fig. 7(a) is a macroscopic view of the fractured surface of UFG410 alloy values, where the striations are > 10.1–0.2 μm, corresponds with
that was tested at 265 MPa. The arrows indicate the origin points of macroscopic crack growth.

6
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 7. SEM image of failed surface of Al5024


UFG410 alloy tested at 265 MPa. (a) Macroscopic
view of the fractured surface where crack initia-
tion sites have been indicated by arrows, (b) re-
gion 1 at high magnification indicating stable
crack growth region, (c) final fracture region 3
and adjacent fatigue striation of different spacing,
and (d) magnified view of fatigue striation.

Fig. 8(a) is the macroscopic view of fatigue failed surface of 5024 Al compared to AA2054 UFG410 alloy in Fig. 7(a), which indicates that
UFG210 alloy tested at 295 MPa. Fig. 8(a) reveals that in region 1, crack 5024 Al UFG210 alloy has slower crack propagation rate than 5024 Al
initiation and stable crack growth rate occupied more area fraction as UFG410 alloy. Mughrabi et al. [16] and Mughrabi [17] mentioned that

Fig. 8. SEM image of failed surface of Al5024


UFG210 alloy tested at 295 MPa. (a) Macroscopic
view of the fractured surface where crack initia-
tion sites have been indicated by arrows, (b) re-
gion 1 at high magnification indication stable
crack growth region, (c) final fracture region and
adjacent fatigue striation of different spacing and
(d) magnified view of fatigue striation.

7
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Fig. 9. SEM image of failed surface of Al5024 FG


alloy tested at 130 MPa (a) Macroscopic view of
the fractured surface where crack initiation sites
are indicated by arrows, (b) region 1 in high
magnification showing faint fatigue striation, (c)
high magnification view of clearly visible fatigue
striation and final overload region and (d) mag-
nified view of fatigue striation.

Table 3 this region (Fig. 9(b)). Average distance between striations was around
Summary of fractographic observations and conclusion. 1 μm (Fig. 9(d)).
The two observations above validated two important conclusions:
Conditions Summary
(i) stage II of crack propagation starts very early in 5024 Al FG alloy as
UFG410 • Various crack initiation points were observed compared to the other two UFG alloys. (ii) Crack propagation rate is
• Crack initiation time was higher, while crack propagation time is
lower than the FG alloy
considerably less in 5024 Al FG alloy as compared to UFG alloys. Slow
crack propagation rate combined with excellent properties of fine grain
UFG210 • Stage I of crack growth occupies more area compared to the
same in both UFG and FG
microstructure enable the 5024 Al FG alloy to exhibit excellent fatigue
properties when compared to normalized stress vs life graph in
• Crack initiation time is higher than UFG and FG alloys
410

410 Fig. 5(b). The above fractographic conclusions are summarized in


FG • No clear distinction between fractographic surface of Stage I
and Stage II of crack propagation is evident
Table 3.

• Stage
UFG
II of crack propagation occurs very early as compared to
210 and UFG 410
4. Conclusions

In this study, two UFG and one fine grain microstructure of 5024 Al
resistance to crack initiation increases with decrease in the grain size. alloy were investigated to understand fatigue properties. EBSD analysis
This will also lead to higher crack resistance in UFG210. A high mag- of interrupted specimen verified that no major grain coarsening took
nification image of this region as presented in 8(b) was very similar to place in 5024 Al UFG alloy, and the microstructure obtained was stable.
Fig. 7(b), and denotes bimodal structure with ridges and plateaus. An increase in intragranular dislocation density using KAM analysis
Fig. 8(c), the higher magnification view of regions 2 and 3 combined, indicated excellent cyclic stability. S-N curves for UFG410 and UFG210
consists of striation marks and final failure dimples. Fatigue striations revealed that within the UFG regime, UFG210 condition with smaller
in 5024 Al UFG210 alloys were less prominent than 5024 Al UFG410 grains has better fatigue properties on the basis of absolute stress levels;
alloy (Fig. 7(c) and (d)). If plastic blunting/re-sharpening model for while in normalized stress, UFG210 shows improvement in high cycle
striation formation by Laird [37] is considered, in ductile face-centered fatigue region and UFG410 in low cyclic region. Basquin equation
cubic materials, duplex slip is the main criterion for striation formation. parameter analysis of all the three microstructure along with data
Since decrease in grain size portends an increased difficulty of dis- points from literature points to the microstructure stability of FSP
location movement leading to a smaller planar slip, striation formation material. Fractography study confirmed that crack initiation and stable
becomes less difficult [38]. crack propagation rate increased with reduction in grain size; while fine
Average striation spacing in 5024 Al UFG210 alloy close to 1.7 μm grain microstructure had excellent crack propagation properties,
and almost the same as 5024 Al UFG410 alloy implies that the crack making it comparable in normalized stress vs life curve.
propagation rate is approximately similar in both alloys. All the above
observation indicated that UFG210 has a higher crack resistance than Acknowledgment
UFG410. Fig. 9(a) shows the low magnification image of 5024 Al FG
failed sample tested at 135 MPa. The magnified view of region 1 is very The authors gratefully acknowledge the support of National Science
close to crack initiation, and faint striation marks can be observed in Foundation (NSF) through grant no 1435810. Authors also thank the

8
S. Shukla et al. International Journal of Fatigue 109 (2018) 1–9

Materials Research facility (MRF) at University of North Texas for the alloys. Metall Mater Trans A 1990;21:421–30.
use of microscopy facility. [19] Kendig KL, Miracle DB. Strengthening mechanisms of an Al-Mg-Sc-Zr alloy. Acta
Mater 2002;50:4165–75.
[20] Komarasamy M, Mishra RS. Serration behavior and shear band characteristics
References during tensile deformation of an ultrafine-grained 5024 Al alloy. Mater Sci Eng, A
2014;616:189–95.
[21] Hall EO. Proc Phys Soc Sect B 1951;64:747.
[1] Liu R, Tian Y, Zhang Z, An X, Zhang P, Zhang Z. Exceptional high fatigue strength in
[22] Kumar N, Mishra RS. Thermal stability of friction stir processed ultrafine grained Al
Cu-15 at. % Al alloy with moderate grain size. Sci Rep 2016;6:27433.
Mg Sc alloy. Mater Charact 2012;74:1–10.
[2] Fleck NA, Kang KJ, Ashby MF. Overview no. 112: the cyclic properties of en-
[23] De PS, Obermark CM, Mishra RS. Development of a reversible bending fatigue test
gineering materials. Acta Metall Mater 1994;42:365–81.
bed to evaluate bulk properties using sub-size specimens. J Test Eval 2008;36:1–4.
[3] Valiev RZ, Islamgaliev RK, Alexandrov IV. Bulk nanostructured materials from se-
[24] Landgraf RW. The resistance of metals to cyclic deformation, Achievement of high
vere plastic deformation. Prog Mater Sci 2000;45:103–89.
fatigue resistance in metals and alloys. ASTM International; 1970.
[4] Saito Y, Utsunomiya H, Tsuji N, Sakai T. Novel ultra-high straining process for bulk
[25] Morrow J. Cyclic plastic strain energy and fatigue of metals. Internal friction,
materials—development of the accumulative roll-bonding (ARB) process. Acta
damping, and cyclic plasticity ASTM International; 1965.
Mater 1999;47:579–83.
[26] Ulacia I, Dudamell NV, Glvez F, Yi S, Prez-Prado MT, Hurtado I. Mechanical be-
[5] Zehetbauer MJ, Valiev RZ. Nanomaterials by severe plastic deformation. John Wiley
havior and microstructural evolution of a Mg AZ31 sheet at dynamic strain rates.
& Sons; 2006.
Acta Mater 2010;58:2988–98.
[6] Witkin DB, Lavernia EJ. Synthesis and mechanical behavior of nanostructured
[27] Yi S, Schestakow I, Zaefferer S. Twinning-related microstructural evolution during
materials via cryomilling. Prog Mater Sci 2006;51:1–60.
hot rolling and subsequent annealing of pure magnesium. Mater Sci Eng, A
[7] Youssef KM, Scattergood RO, Murty KL, Koch CC. Nanocrystalline Al–Mg alloy with
2009;516:58–64.
ultrahigh strength and good ductility. Scr Mater 2006;54:251–6.
[28] Calcagnotto M, Ponge D, Demir E, Raabe D. Orientation gradients and geome-
[8] Huang Y, Langdon TG. Advances in ultrafine-grained materials. Mater Today
trically necessary dislocations in ultrafine grained dual-phase steels studied by 2D
2013;16:85–93.
and 3D EBSD. Mater Sci Eng, A 2010;527:2738–46.
[9] Mishra RS, Ma ZY. Friction stir welding and processing. Mater Sci Eng: R: Rep
[29] Vinogradov A, Washikita A, Kitagawa K, Kopylov VI. Fatigue life of fine-grain
2005;50:1–78.
Al–Mg–Sc alloys produced by equal-channel angular pressing. Mater Sci Eng, A
[10] Valiev RZ, Zhilyaev AP, Langdon TG. Bulk nanostructured materials: fundamentals
2003;349:318–26.
and applications. John Wiley & Sons; 2013.
[30] Kun F, Carmona HA, Andrade Jr JS, Herrmann HJ. Universality behind Basquin’s
[11] Kumar N, Mishra RS, Huskamp CS, Sankaran KK. The effect of friction stir pro-
law of fatigue. Phys Rev Lett 2008;100:094301.
cessing on the microstructure and mechanical properties of equal channel angular
[31] A. Handbook, volume 12: fractography. ASM International Materials Park 1987;55.
pressed 5052Al alloy sheet. J Mater Sci 2011;46:5527–33.
[32] Williams JJ, Yazzie KE, Phillips NC, Chawla N, Xiao X, De Carlo F, et al. On the
[12] Horita Z, Furukawa M, Nemoto M, Valiev RZ, Langdon TG. Characterization of
correlation between fatigue striation spacing and crack growth rate: a three-di-
ultrafine-grained structures produced by severe plastic deformation, Investigations
mensional (3-D) X-ray synchrotron tomography study. Metall Mater Trans A
and applications of severe plastic deformation. Springer; 2000. p. 155–62.
2011;42:3845–8.
[13] Qu S, An XH, Yang HJ, Huang CX, Yang G, Zang QS, et al. Microstructural evolution
[33] Pelloux RM, Faral M, McGee WM. Fractographic measurements of crack-tip closure.
and mechanical properties of Cu–Al alloys subjected to equal channel angular
Fracture mechanics ASTM International; 1980.
pressing. Acta Mater 2009;57:1586–601.
[34] Wanhill R, Hattenberg T. Fractography-based estimation of fatigue crack “initia-
[14] Kamikawa N, Huang X, Tsuji N, Hansen N. Strengthening mechanisms in nanos-
tion” and growth lives in aircraft components; 2006.
tructured high-purity aluminium deformed to high strain and annealed. Acta Mater
[35] Barter SA, Molent L, Wanhill R. Marker loads for quantitative fractography of fa-
2009;57:4198–208.
tigue cracks in aerospace alloys. ICAF 2009, bridging the gap between theory and
[15] Kumar N, Mishra RS, Huskamp CS, Sankaran KK. Critical grain size for change in
operational practice. Springer; 2009. p. 15–54.
deformation behavior in ultrafine grained Al–Mg–Sc alloy. Scr Mater
[36] Lynch S. Some fractographic contributions to understanding fatigue crack growth.
2011;64:576–9.
Int J Fatigue 2017;104:12–26.
[16] Mughrabi H, Höppel HW. Cyclic deformation and fatigue properties of very fine-
[37] Laird C. The influence of metallurgical structure on the mechanisms of fatigue crack
grained metals and alloys. Int J Fatigue 2010;32:1413–27.
propagation. Fatigue crack propagation ASTM International; 1967.
[17] Mughrabi H. On the grain-size dependence of metal fatigue: outlook on the fatigue
[38] De PS, Mishra RS, Smith CB. Effect of microstructure on fatigue life and fracture
of ultrafine-grained metals. Investigations and applications of severe plastic de-
morphology in an aluminum alloy. Scr Mater 2009;60:500–3.
formation Springer; 2000. p. 241–53.
[18] Sawtell RR, Jensen CL. Mechanical properties and microstructures of Al-Mg-Sc

You might also like