Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

JMEPEG (2020) 29:4872–4878 ÓASM International

https://doi.org/10.1007/s11665-020-04696-y 1059-9495/$19.00

Microstructure, Thermal Stability during Creep


and Fractography Study of Friction-Stir-Processed
AA2024-T3 Aluminum Alloy
Michael Regev and Stefano Spigarelli

(Submitted September 19, 2019; in revised form February 18, 2020; published online February 27, 2020)

Friction stir processing (FSP) makes it possible to obtain a stir zone with very fine grain size with the aid of
severe plastic deformation. Yet using FSP, it is impossible to obtain a uniform cross section as far as the
microstructure and mechanical properties are concerned. To reduce the effect of this limitation, in the
current study, the material was processed on both sides, thus yielding a wider, rectangular and more
homogenous stir zone. In a recent publication, the authors focused on the mechanical properties, thermal
stability and transmission electron microscopy (TEM) study of friction-stir-processed AA2024-T3, com-
paring this alloy to the parent material. While the previous study mentioned focused on the parent and on
the as friction-stir-processed material, the current study focuses on post-creep specimens and hence com-
pletes the previous one with microstructural processes occurring during creep. The current paper com-
pletes the above precipitate analysis using electron-dispersive x-ray spectroscopy (EDS) mapping of the
various precipitates reported after exposure to creep temperatures. The TEM study reported in the current
paper revealed the formation of dislocation structures during creep, in parallel to the dynamic recrystal-
lization (DRX) reported by the authors in the past. In addition, fractography study indicated premature
cracking as the prevailing failure mechanism as well as in the case of friction stir welded creep specimens.

et al. (Ref 4) focused on mechanical properties of FSP AA2024.


Keywords AA2024-T3, aluminum, creep, friction stir processing
Cavaliere (Ref 5) also reported on the superplastic behavior of
friction-stir-processed 2024 Al alloy and further studied the
effect of the addition of Sc and Zr, reporting on the average
grain size of 1 lm at the stir zone. Nadammal et al. (Ref 6)
studied microstructure and texture evolution during single- and
1. Introduction multiple-pass FSP. They reported on the formation of equiaxed
grains with an average size of 4-5 lm at the stir zone, claiming
In 2000, FSP was derived from friction stir welding (FSW) that dynamic recrystallization (DRX) was the dominant mech-
with the aim of using severe plastic deformation to obtain a stir anism, while particle stimulated nucleation (PSN) was a
zone with very fine grain size (Ref 1, 2). FSP is identical to participating nucleation mechanism. Ren et al. (Ref 7) studied
FSW except that in FSP the rotating tool does not weld the parts crack repairing by FSP in the 2024 aluminum alloy. They used
to one another. Thus, its operation may be referred to as a optical microscopy for their microstructure study and provided
‘‘bead on plate’’ process. Very few publications deal with FSP microhardness profiles, but they did not conduct any quanti-
of the 2024 aluminum alloy (Ref 2-10). Nadammal et al. (Ref tative analysis of the microstructure at the repaired zone.
2) applied a bottom-up approach for optimizing the process Nevertheless, they did indicate that DRX was the process
parameters; their results included mechanical properties and taking place underneath the shoulder of the tool. Ghanbari et al.
residual stress analysis together with grain size and precipitate (Ref 8) reported on single- and multi-pass FSP of AL2024/SiC
studies conducted using scanning electron microscopy (SEM). composite. They noted that optical microscopy revealed fine
Charit and Mishra (Ref 3) studied the superplastic behavior of equiaxed grains at the stir zone but did not report on the size of
friction-stir-processed AA2024-T4. Based on qualitative trans- these grains. Also, they referred to the influence of the number
mission electron microscope (TEM) examination, they claimed of passes, between 1 and 4, on the hardness of the processed
that the grain size obtained at the stir zone was 3.9 lm. Suri material, noting a decrease in hardness as the number of passes
increased. El-Mahallawi et al. (Ref 9) reported on an increase in
hardness, from 94 HV in the case of the non-processed material
This article is an invited submission to JMEP selected from to 114.7 for material after undergoing FSP together with an
presentations at the Symposium ‘‘Joining and Related increase in the UTS. Nadammal et al. (Ref 10) studied
Technologies,’’ belonging to the topic ‘‘Processing’’ at the European
Congress and Exhibition on Advanced Materials and Processes microstructure evolution during FSP of AA2024, AA2219 and
(EUROMAT 2019), held September 1-5, 2019, in Stockholm, AA5086 using x-ray diffraction and electron backscattered
Sweden, and has been expanded from the original presentation. diffraction techniques. They emphasized the role of strain-
induced grain boundary migration (SIMB) and particle-stimu-
Michael Regev, Department of Mechanical Engineering, ORT Braude lated nucleation (PSN) that acted simultaneously as the
College, P.O. Box 78, Karmiel 21982, Israel; and Stefano Spigarelli, nucleation mechanisms for DRX in Al alloys during FSP,
DIISM, Università Politecnica delle Marche, Ancona 60131, Italy.
Contact e-mail: michaelr@braude.ac.il. while in the case of AA2024 PSN, large particles of Al2CuMg

4872—Volume 29(8) August 2020 Journal of Materials Engineering and Performance


became nuclei for DRX. It seems that none of these publica-
tions conducted an analytical transmission electron microscopy
(TEM) study of the friction-stir-processed material, nor did they
study its aging behavior. Moreover, only two of them (Ref 6, 8)
dealt with multiple pass and its effect on the microstructure and
mechanical properties of the stir zone.
The stir zone of a friction-stir-processed material is as wide
as the diameter of the shoulder at the upper surface of the weld
and markedly narrower near its opposite surface. This property
as well as the differences between the advancing side and the
retreating side makes it impossible to obtain a uniform cross
section as far as the microstructure and mechanical properties
are concerned. For these reasons, in a recent publication, the
authors offered a new approach in which the material was
processed on both sides, thus yielding a wider, rectangular and
more homogenous stir zone (Ref 11) from which all the
specimens were machined out. Processing the material from
both sides eliminated any microstructural difference between
the upper and the lower side at least within the gage lengthÕs
cross section of the creep specimens. The authorsÕ aforemen-
Fig. 1 The FSP tool
tioned recent publication (Ref 11) focused on the mechanical
properties of friction-stir-processed material, on the materialÕs
thermal stability and on transmission electron microscopy
(TEM) studies of friction-stir-processed AA2024-T3 that
compared it to the parent material. As stated earlier, contrary
to the previous study (Ref 11) that dealt with the parent and
with the as friction-stir-processed material, the current study
focuses on post-creep specimens and hence completes the
previous one with microstructural processes occurring during
creep. The current paper completes the previous TEM study
with an investigation of friction-stir-processed AA2024-T3
after creep. The TEM study reported in the current paper
revealed the formation of dislocation structures during creep, in Fig. 2 The different zones obtained during a single-pass FSP
parallel to the dynamic recrystallization (DRX) reported in (Ref
11). The current paper adds precipitate analysis using electron- flipped and processed once again right above the first pass,
dispersive x-ray spectroscopy (EDS) mapping of the various which was on the bottom side of the plate. The second pass was
precipitates after exposure to creep temperatures, hence relay- made so that the advancing side of the first pass became the
ing on a markedly broader statistics of precipitates than the retreating side of the second and vice versa. After trial and
previous publication (Ref 11). In addition, this paper presents error, the optimal welding parameters were found to be a
and discusses a fractography study of broken creep specimens. rotational speed of 800 rpm and a transverse speed of 80 mm/
min.
Grain size measurement was taken using Olympus Stream
Essentials image analysis software. Vickers microhardness
2. Experimental Procedure measurements were taken using a Seiki Matsuzawa microhard-
ness tester under a load of 200 gf. Creep specimens, having a
The material used for this study was commercial AA2024- 3 mm 9 3 mm square cross section and 25 mm gage length,
T3 aluminum alloy in the form of 200 mm 9 100 mm plates, were machined from the FSP-region of the plates. The
3.175 mm thick. The above plates were subjected to FSP using dimensions and the shape of these specimens were the same
a SHARNOA CNC milling machine. The H-13 steel welding/ as of those used in previous studies (Ref 11-14). The
processing tool used consisted of a conical pin with 4.5 mm top longitudinal axis of the samples was parallel to the FSP
diameter and 3 mm in height and a 20 mm diameter shoulder, direction, so that the whole body of the specimen was included
as shown in Fig. 1. in the stir zone. A three-zone furnace was used for heating the
Figure 2 depicts a general illustration showing the different samples up to testing temperature. Constant load creep
zones obtained during FSP. The only zone relevant for the experiments (CLE) were carried out on the samples at 250
current study was the stir zone. From this zone, the creep and and 315 °C under 30-140 and 12-60 MPa, respectively, in most
the tensile specimens were machined out, with their longitu- cases up to sample fracture. Additional variable load experi-
dinal axis parallel to the longitudinal axis of that zone. In ments (VLE) were carried out at the same temperature.
addition, the stir zone of a friction-stir-processed material is as Additional details regarding the creep tests were reported
wide as the diameter of the shoulder at the upper surface of the elsewhere (Ref 13). The TEM investigation was conducted
weld and markedly narrower near its opposite surface. In order using an FEI Tecnai G2 T20 TEM, while EDS mapping was
to make the stir zone as symmetrical and uniform as possible generated with the aid of an FEI Titan Themis G2 60-300 High-
while achieving an almost rectangular cross section, a single Resolution TEM (HRTEM). The fractography study was
pass was made on one side of the plate. The plate was then conducted on broken friction-stir-processed creep specimens

Journal of Materials Engineering and Performance Volume 29(8) August 2020—4873


with the aid of an FEI Inspect SEM. TEM specimens were grain boundary-decorating precipitates as well as the rod-
prepared using an FEI Helios NanoLab G3 FIB (focused ion shaped and nano-sized precipitates reported in (Ref 11) are all
beam). discernible in Fig. 5(a) and marked as precipitates 1, 2 and 3,
respectively. The region of interest (ROI) on which EDS
mapping was carried out as well as the different types of
precipitates is shown in Fig. 5(a). A DF image of the ROI is
3. Results shown in Fig. 5(b), while Fig. 5(c) and (d) depicts qualitative
Cu and Mg content, respectively. Note that all the precipitates
As stated earlier, in a recent publication (Ref 11), the authors appearing in Fig. 5(b) contain Cu, while some of them contain
reported the tensile and microhardness properties of the Mg and others do not.
friction-stir-processed material compared to those of the parent Figure 6(a) is a bright field (BF) TEM micrograph of a
material. The main findings indicated that the processed specimen that crept for 53 min under 25 MPa at 315 °C, taken
material showed inferior tensile properties relative to the parent near <011> zone axis (z.a.), showing the dislocation structure
material. These inferior tensile properties together with their comprising sub-grain walls next to grain boundary precipitates.
wide scattering are related to defects introduced during the FSP. These dislocation networks can be clearly seen under higher
The coarse elongated grains of the parent material changed into magnification in Fig. 6(b), while Fig. 6(c) shows the respective
fine equiaxed ones, and very coarse precipitates tens of microns selected area electron diffraction patterns (SADP).
in size detected at the parent material were broken into Figure 7(a) provides a general view of the fracture surface
uniformly dispersed 0.1-1 lm precipitates. However, optical of a friction-stir-processed specimen that crept to rupture under
metallography of the stir zone under higher magnifications 40 MPa at 315 °C. Three different zones are depicted in
(Fig. 3) revealed that the grain size varies by more than an Fig. 7(a) and shown under higher magnification in Fig. 7(b)-
order of magnitude. The average grain size was found to be 5.1 (d). The right side of the fracture surface (zone 1 in Fig. 7a and
microns, the coarsest grains were found to be about 11 microns b) is a ductile oxidized surface. The surface of the central zone
in diameter, while the finest grains were found to be about two (zone 2 in Fig. 7a and c) looks like a strongly oxidized ductile
microns in diameter according to the image analysis software fracture surface. Typical dimples are shown in Fig. 7(b) and (c).
and less than one micron according to manual mean intercept As for zone 3 in Fig. 7(a) (and Fig. 7d), besides having a more
length measurements. brittle character, it also contains some striation-like marks that
Figure 4 shows three microhardness profiles taken from the can be clearly observed in Fig. 7(d). The arrow shown in
stir zone. Keeping in mind that the width of the gage length of Fig. 7(e) points at a crack. This crack is assumed to be the
the tensile/creep specimens was 3 mm, one can see that these initiation point of the fracture, i.e., the fracture started at zone 1
profiles demonstrate the hardness of the specimens. Profile 1 and propagated to zone 2, while the cracking in zone 3 occurred
was taken as close as possible to the upper surface, profile 2 at the end of the process.
was taken at the middle of the cross section, and profile 3 was
taken near the lower surface. All the measurements were taken
across the stir zone, namely perpendicular to its longitudinal
axis, while maintaining a distance of 0.25 mm from one 4. Discussion
indentation to another. Although the hardness of the stir zone is
higher than that measured in the case of the parent material, i.e., One of the conclusions of the first part of this study (Ref 11)
about 100 HV (Ref 11), a certain degree of hardness variation was that the FSP conducted on both sides of the 3.125-mm-
is discernible. Figure 4 shows that the hardness values of the thick AA2024-T3 plates yielded inferior tensile properties
stir zone vary from 108 to 130 HV. relative to the parent material . The wide scattering of the
Figure 5(a) depicts a high-angle annular dark-field scanning tensile results leads to the conclusion that defects introduced
transmission electron microscope (HAADF-STEM) image of a during the FSP were responsible for the poor tensile properties.
specimen that crept for 53 min under 25 MPa at 315 °C. The The hardness variation inside the stir zone along each profile as
shown in Fig. 4 (the hardness of profile 1 varies from 121.2 to
129.9 HV, the hardness of profile 2 varies from 111.1 to
120.7 HV, and that of profile 3 varies from 107.8 to 120.2 HV)
can be related to the grain size difference observed in Fig. 3. As
stated earlier, the grain size varied by more than an order of
magnitude, from less than one micron to more than ten microns.
Keeping in mind the Hall–Petch dependence of strength on the
inverse square root of grain size, the above hardness variation
can be attributed to the differences in the grain size observed.
Consequently, further technological effort is still required in
order to obtain a completely homogeneous stir zone.
In the current study, also the creep properties were found to
be inferior, compared to those of the base material, namely in
its as-produced condition. For example, in the case of 315 °C
and 40 MPa, the minimum creep rate recorded for the base
material was 5.0 9 10 7 s 1 and the time to reach this
Fig. 3 A representative optical micrograph of the stir zone minimum creep rate was 8 h (Ref 14), while in the case of

4874—Volume 29(8) August 2020 Journal of Materials Engineering and Performance


Fig. 4 Microhardness profiles across the stir zone: 1—taken near the upper surface; 2—taken at the middle of the cross section; 3—taken near
the lower surface

Fig. 5 EDS mapping of a specimen that crept 53 min under 25 MPa at 315 °C: (a) HAADF-STEM micrograph showing 1—grain boundary
decoration, 2—rod-shaped precipitates, 3—nano-sized precipitates; (b) HAADF-STEM micrograph of the ROI; (c) qualitative EDS mapping
showing Cu concentration; (d) qualitative EDS mapping showing Mg concentration

the friction-stir-processed material the minimum creep rate was 1. The high stress exponent n in the power law describing
1.0 9 10 5 s 1 and the whole test, until fracture occurred, took the minimum creep rate dependence on applied stress,
0.42 h. In principle, the higher creep rates of the fine-grained which is well above the n = 2 value typical of GBS-con-
FSW material could be attributed to the occurrence of grain trolled creep (Ref 13).
boundary sliding (GBS). Yet, there are several reasons to rule
out any significant role of GBS, namely:

Journal of Materials Engineering and Performance Volume 29(8) August 2020—4875


On these bases, one can safely conclude that also in the FSP
alloy, creep is controlled by dislocation motion, and that the
precipitate size and distribution play a significant role.
In previous studies (Ref 11, 12), the authors found that the
microstructure of the friction-stir-processed material (Ref 11) is
unstable, as is that of the parent AA2024 (Ref 12). Contrary to
the single precipitate analyses in the previous studies, the
systematic EDS mappings conducted on the friction-stir-
processed material during the current study unambiguously
indicated that the precipitates formed during creep can contain
either Cu, Mg and Al or just Cu and Al (see Fig. 5). This is in
line with the precipitate study of AA2024, which many
researchers continue to pursue and according to which two
different aging sequences take place—the h sequence in which
Al2Cu precipitates are formed and the S sequence that is
responsible for the Al2CuMg (Ref 11).
The TEM study reported in (Ref 11) revealed a low-angle
sub-grain structure with dislocation network boundaries in the
case of the parent metal. During FSP, these were replaced with
fine equiaxed grains having clean boundaries with no disloca-
tion tangles. As can be seen in Fig. 6, dislocation networks
appeared in the friction-stir-processed material after creep
appeared at the grains that underwent dynamic recrystallization
(DRX) during the FSP, namely the recrystallized grains with the
clean boundaries mentioned above. The authors previously
proposed (Ref 11) that the high stacking fault energy of Al,
which is responsible for dynamic recovery (DRV) during the
thermo-mechanical processing of the 2024 aluminum alloy
(i.e., rolling in the case of the parent material), leads in turn to
the arrangement of dislocations into sub-grain boundaries,
while the high strain rates of the FSP lead to DRX. Hence, it
may be claimed that the low creep strain rates, which are even
lower than those of the rolling process, are responsible for the
reappearance of the dislocation tangles in the crept friction-stir-
processed specimen. The material was also found to undergo
DRX during the creep process (Ref 12). Thus, it appears that
both DRX, which is responsible for ultra-fine grain formation,
and sub-grain formation act during creep. Further research,
however, is still required in order to clarify whether these two
mechanisms act simultaneously or one after another. One of the
possibilities is that DRX occurs at the beginning of the process,
while sub-grain formation takes place at a later stage. This can
be checked by conducting creep tests and interrupting them
after reaching various amounts of creep strain combined with
TEM study on the respective creep specimens.
In the current study, the creep results were found to be
inferior to those of the base material, namely in its as-produced
condition. For example, in the case of 315 °C and 40 MPa, the
minimum creep rate recorded for the base material was
5.0 9 10 7 s 1 and the time to reach this minimum creep rate
was 8 h (Ref 14), while in the case of the friction-stir-processed
material the minimum creep rate was 1.0 9 10 5 and the entire
Fig. 6 TEM micrographs of specimen that crept for 53 min under test until fracture took 0.42 h. In a previous study that dealt
25 MPa at 315 °C showing sub-grain boundaries taken near with FSW of AA2024-T3 (Ref 12), the authors found that an
<011> z.a.; (a, b) BF images under different magnifications; (c) open-to-surface cracking limited the creep resistance and
SADP of <011> z.a discussed its influence in detail. Figure 7(e) shows an identical
failure mechanism. Namely, failure starts with the formation of
2. The presence of an almost continuous chain of intergran- an open-to-surface crack, which then propagates into the
ular precipitates (Fig. 5), which can effectively obstruct material. The existence of this crack markedly limits the creep
any sliding along grain boundaries. resistance and leads to premature failure. It may be concluded,
3. The formation of a substructure during creep, typical of therefore, that eliminating this cracking is essential to improve
climb-controlled dislocation motion. creep resistance, though further research is required to corrob-
orate this conclusion. As for the striation-like phenomenon

4876—Volume 29(8) August 2020 Journal of Materials Engineering and Performance


Fig. 7 Fracture surface of a friction-stir-processed specimen that crept to rupture under 40 MPa at 315 °C: (a) general view; (b) zone 1 under
high magnification; (c) zone 2 under high magnification; (d) zone 3 under high magnification; (e) crack initiation

detected in zone 3 (see Fig. 7d), this is due to plastic tate statistics, that both the h and the S sequences take
deformation bands resulting from the heavy plastic deformation place during exposure to creep temperature
during FSP. 3. The TEM study provided evidence for sub-grain forma-
tion during creep inside the grains that had undergone
DRX during the FSP. This in turn indicates that, in addi-
tion to precipitation processes, both DRX and sub-grain
5. Conclusions formation occur during creep. However, further research
is still required in order to clarify whether these two
1. AA2024-T3 underwent friction stir processing on both
mechanisms act simultaneously or one after another.
sides, thus yielding a wider, rectangular and more
4. Cracking was found to take place during creep of the
homogenous stir zone.
friction-stir-processed material as well as in the case of
2. Precipitate analysis conducted by using EDS mapping
the friction stir welded material, hence reducing creep
provided evidence, which relied on quite broad precipi-
resistance.

Journal of Materials Engineering and Performance Volume 29(8) August 2020—4877


5. Further technological effort is required in order to elimi- 6. N. Nadammal, S.V. Kailas, J. Szpunar, and S. Suwas, Microstructure
nate the cracking and improve the creep resistance of the and Texture Evolution during Single and Multiple-Pass Friction Stir
Processing of Heat-Treatable Aluminum Alloy 2024, Metall. Mater.
friction-stir-processed AA2024-T3 as well as to obtain a Trans. A, 2017, 48(9), p 4247–4261
completely homogeneous stir zone with a uniform grain 7. J.G. Ren, L. Wang, D.K. Xu, L.Y. Xie, and Z.C. Zhang, Analysis and
size. Such a possible solution can be multi-pass process- Modeling of Friction Stir Processing-Based Crack Repairing in 2024
ing on each side of the material. Aluminum Alloy, Acta Metall. Sin. (Engl. Lett.), 2017, 30(3), p 228–
237
8. D. Ghanbari, M. Kasiri Asgharani, and K. Amini, Investigating the
Effect of Passes Number on Microstructural and Mechanical Properties
of the Al2024/SiC Composite Produced by Friction Stir Processing,
Acknowledgments Mechanika, 2015, 21(6), p 430–436
9. I. El-Mahallawi, M.M.Z. Ahmed, A.A. Mahdy, A.M.M. Abdelmota-
The assistance of Dr. Alexander Katz-Demyanetz with the galy, W. Hoziefa, and M. Refat, Effect of Heat Treatment on Friction-
fractography study is highly appreciated. The authors thank Dr. Y. Stir-Processed Nanodispersed AA7075 and 2024 Al Alloys, Friction
Kauffmann for his assistance with the TEM study and EDS Stir Welding and Processing IX, Friction Stir Welding and Processing
IX, The Minerals, Metals & Materials Society, 1st ed., Y. Hovanski, R.
mapping and Dr. L. Popilevsky for TEM specimen preparation. Mishra, Y. Sato, P. Upadhyay, and D. Yan, Ed., Springer International
Publishing, San Diego, 2017, p 297–309
10. N. Nadammal, S.V. Kailas, J. Szpunar, and S. Suwas, Restoration
Mechanisms During the Friction Stir Processing, Metall. Mater. Trans.
Funding A, 2015, 46(7), p 2823–2828
This research project is partially funded by Ort Braude College. 11. M. Regev and S. Spigarelli, Study of Mechanical, Microstructural and
Thermal Stability Properties of Friction Stir Processed Aluminum
2024-T3 Alloy, Kovove Mater., 2019, 57(4), p 229–236
12. M. Regev, T. Rashkovsky, M. Cabibbo, and S. Spigarelli, Microstruc-
References ture Stability During Creep of Friction Stir Welded AA2024-T3 Alloy,
J. Mater. Eng. Perform., 2018, 27(10), p 5054–5063
1. R.S. Mishra, M.W. Mahoney, S.X. McFadden, N.A. Mara, and A.K. 13. E. Santecchia, M. Cabibbo, M. Ghat, M. Regev, and S. Spigarelli,
Mukherjee, High Strain Rate Superplasticity in a Friction Stir Physical Modeling of the Creep Response of an Al-Cu-Mg Alloy With
Processed 7070 Al Alloy, Scr. Mater., 2000, 42(2), p 163–168 a Fine Microstructure Transformed by Friction Stir Processing, Mater.
2. N. Nadammal, S.V. Kailas, and S. Suwash, A Bottom-Up Approach for Sci. Eng. A, 2020, 769, p Article 138521
Optimization of Friction Stir Processing Parameters; a Study on 14. C. Paoletti, M. Regev, and S. Spigarelli, Modelling of Creep in Alloys
Aluminium 2024-T3 Alloy, Mater. Des., 2015, 65, p 127–138 Strengthened by Rod-Shaped Particles: Al-Cu-Mg Age-Hardenable
3. I. Charit and R.S. Mishra, High Strain Rate Superplasticity in a Alloys, Metals, 2018, 8(11), p 930-1–930-18
Commercial 2024 Al Alloy Via Friction Stir Processing, Mater. Sci.
Eng. A, 2003, 359(1–2), p 290–296
Publisher’s Note Springer Nature remains neutral with regard to
4. A. Suri, A. Sahai, K.H. Raj, and N.K. Gupta, Impact and Tensile
Testing of Al2024 Alloy Processed by Friction Stir Processing, jurisdictional claims in published maps and institutional affilia-
Procedia Eng., 2017, 173, p 679–685 tions.
5. P. Cavaliere, Effect of Minor Sc and Zr Addition on the Mechanical
Properties of Friction Stir Processed 2024 Aluminium Alloy, J. Mater.
Sci., 2006, 41(13), p 4299–4302

4878—Volume 29(8) August 2020 Journal of Materials Engineering and Performance

You might also like