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

Dissimilar linear friction welding between a SiC particle reinforced

aluminum composite and a monolithic aluminum alloy: Microstructural,


tensile and fatigue properties
F. Rotundo
a,b,n
, A. Marconi
a,c
, A. Morri
c
, A. Ceschini
c
a
Department of Mechanical Engineering (DIEM), Alma Mater Studiorum, University of Bologna, V.le Risorgimento 4, 40136 Bologna, Italy
b
Interdepartmental Center for Industrial Research, Advanced Applications in Mechanical Engineering and Materials Technology (CIRI-MAM), University of Bologna,
V.le Risorgimento 4, 40136 Bologna, Italy
c
Department of Metals Science, Electrochemistry and Chemical Techniques (SMETEC), Alma Mater Studiorum, University of Bologna, V.le Risorgimento 4, 40136 Bologna, Italy
a r t i c l e i n f o
Article history:
Received 15 May 2012
Received in revised form
10 September 2012
Accepted 11 September 2012
Available online 18 September 2012
Keywords:
Metal matrix composite (MMC)
Linear friction welding (LFW)
Dissimilar welding
Mechanical properties
Fatigue
Aluminum
a b s t r a c t
The present work is aimed at evaluating the feasibility of the linear friction welding process to produce
dissimilar joints between a AA2124/25%vol SiC
p
composite and a 2024 Al alloy, illustrating and
correlating their microstructural and mechanical properties. Optical microscopy (OM) and scanning
electron microscopy (SEM) with energy dispersive spectroscopy (EDS) were used to characterize the
effects of the welding process on the microstructure of the LFW joints. Tensile tests were carried out on
joints welded at three different welding parameters. Axial fatigue tests were carried out under stress
control and the corresponding SN probability curves were computed. The mechanisms of failure were
investigated by SEM analyses of the fracture surfaces. In the LFW joints almost no blending between the
MMC and the base Al alloy was detected, while good particle distribution and no clustering were found
on the MMC side. Mechanical testing demonstrated that high quality dissimilar joints, characterized by
good tensile and fatigue properties, with respect to the AA2024 base material, were obtained by means
of LFW. Fracture was usually located in the weld center or in the thermo-mechanically affected zone
(TMAZ), due to the plastic ow which the material underwent during the solid state welding process.
& 2012 Elsevier B.V. All rights reserved.
1. Introduction
Taking advantage of particular properties of the constituent
materials to meet specic demands is the main driver for the
development of metal matrix composites (MMCs). In particular,
when an aluminum alloy is reinforced with ceramic particles, an
increase in specic stiffness and strength, wear resistance and
thermal stability can be achieved, still maintaining the isotropy of
mechanical properties [1,2]. Particle reinforced Al-based MMCs are
used in aerospace and automotive elds, surface transportation
systems, electronic substrates, packaging machines and a variety
of other applications. However, due to the high production costs and
relatively poor impact properties, which can be signicantly lower
than the unreinforced monolithic alloys, the industrial use of MMCs
is particularly attractive in components, or part of components,
where their specic properties provide a signicant performance
increase. Hence, the study of welding methods able to produce high
quality dissimilar joints between Al-based MMCs and unreinforced
Al monolithic alloys is of particular interest.
Major limitations for the application of Al-based MMCs are the
low mechanical properties of the joint obtained with the tradi-
tional fusion welding techniques [3].
In fact, the difculties in welding aerospace grade aluminum
alloys, such as 2XXX (AlCu) and 7XXX (AlZnMg) series, have
hitherto limited their use in structural applications, even in their
monolithic form [4,5]. In particular, the use of traditional fusion
welding techniques generally leads to microstructural defects such
as high porosity, oxide inclusions and solidication shrinkage, as
well as distortions. When reinforced with ceramic particles, further
defects arise due to the presence of the hard phase [610]:
solidication cracks, caused by high thermal stresses at the
matrix/reinforcement interface, due to the differences between
their thermal expansion coefcients;
presence of reinforcement depleted zones or reinforcement
clusters, induced by the poor wettability of the reinforcements
and/or their different density in respect to the matrix;
high level of gas porosity, due to the high viscosity of the melted
composite which does not allow the occluded gas (mainly hydro-
gen) to escape from the welding pool during solidication;
Contents lists available at SciVerse ScienceDirect
journal homepage: www.elsevier.com/locate/msea
Materials Science & Engineering A
0921-5093/$ - see front matter & 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.msea.2012.09.033
n
Corresponding author at: Department of Mechanical Engineering (DIEM), Alma
Mater Studiorum, University of Bologna, V.le Risorgimento 4, 40136 Bologna, Italy.
Tel.: 39 51 2093141; fax: 39 51 2093467.
E-mail address: fabio.rotundo@unibo.it (F. Rotundo).
Materials Science & Engineering A 559 (2013) 852860
possible formation of undesired brittle phases (Al
4
C
3
, MgAl
2
O)
at the matrix/reinforcement interface.
Such defects result in a signicant reduction of tensile and
fatigue strength, ductility and corrosion resistance of the joint.
These effects are even amplied in the production of dissimilar
joints between reinforced and unreinforced alloys, because of
their different thermal, chemical and deformation behaviors.
All these problems can be signicantly reduced by the use of
solid state joining techniques, such as friction welding (FW).
Among these techniques, friction stir welding (FSW) has been
successfully applied to similar and dissimilar Al-based MMCs and
unreinforced Al alloys [1116]. In the case of ceramic reinforced
MMCs, however, one major limitation for this process is the
severe wear of the pin, which possibly leads to Fe entrapment
in the welded zone [11]. This drawback could be overcome by
means of the linear friction welding (LFW) technique [17]. LFW is
a solid state joining process in which a joint between two metals
can be formed through the intimate contact of a plasticized layer
at the interface of the adjoining components. This plasticized
layer is created through a combination of frictional heating and
applied force. The former occurs as a result of pushing one
stationary workpiece against another, that is moving in a linear
reciprocating manner [18].
To the authors knowledge, despite the importance of the subject
to widen MMCs industrial application range, only a limited number
of papers have been published concerning dissimilar welding
between Al-based particle reinforced MMCs and monolithic Al alloys
[1921]. Previous studies were carried out by the authors on similar
MMC to MMC LFW joints [17]. The aim of the present work is to
assess the feasibility of using the Linear Friction Welding process to
produce dissimilar joints between the AA2124/25%vol SiC
p
compo-
site and the AA2024 aerospace grade aluminum wrought alloy, by
investigating and characterizing the microstructural and mechanical
properties of the resulting joints.
2. Materials and methods
2.1. Materials
The composite object of the present study was produced by
a powder metallurgy route, combining a 2124 aluminum alloy
(AA2124) matrix and 25 vol% of ne (o3 mm) silicon carbide
particles (SiC
p
). Further details on the production process were
already given in [17]. The composite was supplied as 15 mm thick
forged plates, while the monolithic alloy, 2024 aluminum alloy
(AA2024), was supplied as extruded rectangular plates. Both
materials were heat treated at the T4 condition (solubilization at
505 1C for 1 h, water quenching and aging at room temperature).
LFW was carried out after 3 weeks from heat treatment, while
mechanical tests were performed after about 3 months from the
welding process. This alloy has nearly the same nominal chemical
composition as the MMC matrix, except for a slight difference in
the maximum allowed Fe and Si content (Table 1).
Al and MMC specimens for LFW were machined by means of
Electric discharge machining (EDM) wire cutting. Since the surface
oxide layer originating from the EDM wire cutting was expected to
Table 1
Nominal chemical compositions (wt%) of the AA2124 MMC matrix and the AA2024.
Cr Cu Fe Mg Mn Si Ti Zn Others Al
AA2124 r0.10 3.804.90 r0.30 1.201.80 0.300.90 r0.20 r0.15 r0.25 r0.15 Balance
AA2024 r0.10 3.804.90 r0.50 1.201.80 0.300.90 r0.50 r0.15 r0.25 r0.15 Balance
Fig. 2. Setup of the LFW process prior to welding: static and oscillatory compo-
nent xed in stainless steel clampings.
Fig. 1. Schematic representation of the linear friction welding process and
reference system.
Table 2
Linear friction welding parameters for dissimilar LFW joints.
Axial force Pressure Frequency Set amplitude Burn-off
(kN) (N/mm
2
) (Hz) (mm) (mm)
70 130 50 72 2
85 157 50 72 2
100 185 50 72 2
MMC side
Al side
Fig. 3. Side view of the dissimilar LFW joint showing the plastically deformed
material expelled as ash, both on the MMC and Al side.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 853
be expelled during the LFW, the specimens were not submitted to
any further machining prior to welding.
2.2. Linear friction welding
Specimens with a 1536 mm
2
cross section and 80 mm
length were Linear Friction Welded at TWI (The Welding Institute,
Cambridge, UK), according to the layout shown in Fig. 1. The set-up
of the specimens, xed in stainless steel clampings prior to
welding, is reported in Fig. 2.
In a rst phase, linear motion of the oscillating component
begins and the two specimens are brought into contact. In this
initial phase, the abrasion of the asperities leads to an increase
of the true contact area and thermal softening of the interface.
Provided the concurrent action of pressure and reciprocating
motion supplies sufcient frictional heating, the subsequent
phase is characterized by the expulsion of large wear particles
and by the expansion of the heat-affected zone into the bulk
material. In the following phase, the combination of the axial
force and material softening at the interface causes a part of this
material to be expelled as ash, and, consequently, a decrease of
the total axial length of the joining specimens. When the imposed
axial shortening (burn-off distance) is achieved, the reciprocating
motion is stopped, although the axial force is maintained to
consolidate the weld, causing further axial shortening.
No post-weld heat treatment was carried out. Welding para-
meters were optimized after some preliminary tests and are
reported in Table 2. A total of 12 dissimilar MMC/Al joints were
produced at different axial force levels. Unless differently specied,
images and data presented in the following paragraphs refer to
joints obtained with an axial force of 85 kN.
2.3. Microstructural characterization
Samples for microstructural investigations were cut along the
yz plane in Fig. 1. The metallographic samples were mechanically
ground, polished (according to ASTM E3 [22]) and chemically
etched with Kellers reagent. The microstructural characterization
was carried out by optical microscopy (OM) under polarized light
and scanning electron microscopy (SEM) with an energy dispersive
spectroscopy (EDS) microprobe. In order to investigate possible
effects of the welding process on particle size distribution, image
analyses were carried out on SEM micrographs.
HAZ
TMAZ TMAZ
Weld centre
HAZ
1 mm
Fig. 4. Optical micrograph of the etched central cross section of the dissimilar
LFW joint between a AA2124/25%vol SiC
p
composite and a AA2024 alloy, showing
the weld center, thermo-mechanically affected zone (TMAZ) and heat affected
zone (HAZ).
Fig. 5. AA2024 alloy base material: (a) OM polarized micrograph which shows the presence of grains elongated in the extrusion direction and intermetallic compounds;
(b) backscattered SEM micrograph which highlights the intermetallic compounds; (c) EDS analysis of CuFe based intermetallics.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 854
2.4. Mechanical testing
Vickers hardness proling was performed by means of a 30 kg
load (HV
30
) on a central line across the joint, parallel to the y-axis, on
the transverse cross-sections used for the microstructural analyses.
Flat dog-bone specimens were machined according to ISO/TTA2
[23] for tensile and fatigue tests, with the main axis perpendicular
to the welded xz plane shown in Fig. 1, having a gauge length of
12 mm. For each welding axial force level, three specimens were
tested, at a strain rate of 10
4
s
1
, on a servo-hydraulic machine
equipped with a 100 kN load cell.
Specimens were polished with abrasive papers until reaching a
nal surface roughness lower than about Ra0.2 mm for fatigue
tests, in order to remove the effects of the EDM machining
Particle-free bands
Intermetallics
Fig. 6. AA2124/25%vol SiC
p
composite, base material: (a) OM micrograph which shows the presence of particles free bands; (b) backscattered SEM micrograph which
highlights the intermetallic compounds.
100 m
100m
20 m
Weld center Weld center
Weld center
Intermetallics
500 m
Fig. 7. Optical micrographs of (a) the overview of the joint; (b) weld center under-polarized light, where it is possible to observe the elongated grain of the TMAZ;
(c) presence of intermetallic compounds at the limit of the weld center; (d, e) high magnication of the interface which shows the low blending level between Al and MMC
materials.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 855
process. Fatigue tests were carried out under stress control on a
servo-hydraulic testing machine, equipped with a 100 kN load
cell. Sinusoidal waveform, at a cyclic frequency of 20 Hz and a
load ratio R0, were used, while the run-out was set at 10
7
cycles. A total of 19 fatigue specimens underwent axial fatigue
tests, with maximum load distributed on 5 levels and load steps
of 20 MPa, ranging from 140 MPa to 240 MPa, i.e. in the 3255%
range of the average joints ultimate tensile strength.
In order to analyze the fatigue data and to determine the SN
probability curves, the maximum likelihood method was used [16,17].
This method allows for the consideration of fatigue tests in which
specimens survive achieving run-out, as opposed to least square
methods, which cannot consider these tests to compute SN curves.
Both tensile and fatigue fracture surfaces were investigated by
SEM to understand the involved failure mechanisms.
3. Results and discussion
3.1. Microstructural analysis
Linear friction joints showed the presence of plastically
deformed MMC and Al alloy ash material, expelled in both
parallel and normal directions in respect to the force-motion
plane, as shown in Fig. 3. Although the plastic ow was slightly
lower on the MMC side, ash material was extruded even from
corners. Indeed, a complete weld penetration was observed in all
the analyzed MMC joints after parameter optimization, together
with clear separation between the reinforced and unreinforced
material with a slight oscillation of the weld line (Fig. 4). The
microstructural analysis did not identify any of the typical fusion
welding defects, neither on the Al-alloys side nor on the Al-based
MMCs side.
Four characteristic zones were observed on etched cross
sections of the LFW joint (Fig. 4), similarly to what was reported
in [17] for similar MMC to MMC LFW joints: weld center, thermo-
mechanically affected zone (TMAZ), heat affected zone (HAZ) and
base material (BM).
Both in the MMC and Al base materials, it is possible to observe
the presence of a deformed microstructure. OM image of the base
Al alloy (Fig. 5a) shows grains elongated in the extrusion direction
and the presence of CuFe based intermetallic compounds, as also
highlighted in the backscattered SEM micrograph of Fig. 5(b) and
EDS analysis of Fig. 5(c). In the MMC base material it is possible to
observe the presence of particle free bands, elongated in the
plastic ow direction induced by the forging process, although
grains could not be resolved due to the very ne microstructure
induced by the powder metallurgy (PM) production process
(Fig. 6a). Backscattered SEM micrographs showed a lower amount
of intermetallic compounds with respect to the Al base material
(Fig. 6b) due both to the ner microstructure and the lower
amount of Fe in the AA2124 alloys in respect to the AA2024 alloy.
In the weld center, the concurrent effect of frictional heating
and severe plastic deformation led to a reduction in the inter-
metallic amount present in the base materials together with a
signicant grain renement, which made single grains not resol-
vable even in the Al alloy (Fig. 7). Several authors [13,15,2426],
in fact, highlighted that friction welding processes of Al alloys can
induce dynamic recrystallization and this effect is enhanced in Al-
MMCs, due to the presence of the reinforcement particles which
act as nuclei for recrystallization. Moreover, in this specic alloy,
plastic deformation probably broke the iron brittle compounds
and distributed the constituents, favouring their partial dissolu-
tion at the temperature reached during the welding process. The
weld center zone is asymmetric in respect to the joining line and
not uniform: its total width ranges from about 100 mm to 160 mm.
Although sound joints were realized, Al and MMC materials are
still clearly distinguishable since almost no particles were trans-
ported to the unreinforced Al alloy side, due to the absence of any
stirring tool (Fig. 7d and e). In the MMC side, the Weld Center
showed a uniform particle distribution, whose size was not
inuenced by the welding process, as highlighted by the SEM
analysis in Fig. 8.
The TMAZ is characterized by brosity induced by the severe
plastic ow. In the MMC side, it is possible to observe thin particle
free zones oriented along the plastic ow (Fig. 9) while, in the Al
side, the same orientation in the distribution of Cu-based inter-
metallic compounds was found, with grains elongated in the
direction of material expulsion as ash (Fig. 10).
The HAZ is not subjected to any plastic deformation but
undergoes thermal cycling due to the high temperature achieved
in the joint during welding (close to 4001 according to the process
modeled in [27]).
3.2. Hardness
Hardness proles, shown in Fig. 11, highlighted discontinuity
at the weld line due to the lack of blending. In the MMC side the
joints showed a minimum hardness of 180 HV between 2 and
3 mm from the weld line, whereas an average of 217 HV was
measured in the base material. From this minimum value, hard-
ness increased up to about 200 HV in the center of the weld and
a
b
Fig. 8. SEM micrograph showing a uniform particle distribution in (a) the weld
center (MMC side) and (b) base material (MMC side).
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 856
reached the base material hardness at about 10 mm from the
center. A similar hardness trend (but clearly with lower values)
was also observed in the Al side. The complexity/uctuation of
the hardness proles is related to the concurrent effects of
frictional heating and severe plastic deformation that lead to
grain renement and possible modication of the coherent or
semicoherent strengthening precipitates [14,15] during the weld-
ing process.
According to [28] the aging sequence for S precipitates
(Al
2
CuMg) is
SSS-CuMg co-clusters-GPB2/S(orthorhombic)-S
where SSS means supersaturated solid solution, and GPB the
GuineerPrestonBagaryatsky zones. The presence of different
strengthening precipitates in the friction welding zone, and their
inuence on the hardness of AlCuMg alloys, was already
reported in [29]. According to this study, also considering for
the MMC which is the effect of the SiC particles on the kinetic of
the precipitation sequence, it is possible to hypothesize that in
the HAZ and TMAZ, the LFW induces CuMg co-clusters dissolu-
tion and, if the temperature is high enough, precipitation and/or
coarsening of the stable S phase (Al
2
CuMg), with a consequent
hardness decrease. In the TMAZ, in particular, the higher tem-
perature leads to the formation of coarser S precipitates, with a
further hardness reduction. The high temperature in the weld
center, instead, dissolved both the CuMg co-clusters, the S
precipitates and the others intermetallic compounds, so that the
post-weld natural aging can lead to the formation of new CuMg
co-clusters. The concurrent effect of grain renement, induced by
dynamic recrystallization, and presence of CuMg co-clusters can
explain the higher hardness in the weld center in respect to the
TMAZ. The hardness uctuations in the joint, especially in the
Fig. 9. Optical micrograph of the TMAZ in the MMC side: (a) overview; (b) presence of thin particle free zones oriented along the plastic ow.
500m
Intermetallics
Fig. 10. Optical micrograph of the TMAZ in the Al side: (a) overview; (b) presence of intermetallic compounds oriented according to the plastic ow.
Fig. 11. Hardness prole across the dissimilar LFW joint between a AA2124/25%vol SiC
p
composite (left) and a AA2024 alloy (right). The weld center is highlighted by the
yellow lines (260 mm thickness). (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 857
TMAZ could be explained by the local strain hardening of severely
deformed material.
3.3. Tensile tests
The tensile tests evidenced that the variation of axial force and,
consequently, of friction pressure, does not seem to have inu-
ence on the mechanical properties, as shown by the experimental
data in Table 3. The proof strength (YS) ranges from 322 to
330 MPa, the ultimate tensile strength (UTS) from 433 to 443 MPa
and the elongation to failure (E%) from 4.7% to 4.8%.
The joint efciency, evaluated as the ratio between the joint
and the nominal AA2124 base material properties [3032], is
higher than 90% for both UTS and YS, but it is about 25% for the
E%. The low elongation to failure is due to the lower ductility and
higher stiffness of the MMC with respect to the unreinforced
alloy, so that the deformation mainly takes place on the Al side.
Moreover, this value is strongly inuenced by the localization of
the failure, as showed by the fracture surfaces analysis.
Tensile fracture surface are characterized by the typical cup-and-
cone shaped morphology. Low magnication SEM analysis of the
fracture surfaces highlighted the presence of at and rough zone in
the central area of the specimens in both the Al and MMC side
(Fig. 12). At higher magnication, it was possible to dene that (i) the
at zones correspond to the fracture in the Al side of the central zone,
with ultra-ne dimples (do1 mm) conrming the grain renement
induced by the LFW; (ii) the rough zones correspond to the fracture in
Table 3
Results of the tensile tests on the LFW joints obtained under different welding
conditions (average values and standard deviations).
Axial force (kN) R
p02
(MPa) UTS (MPa) Elongation to failure %
70 322 SD 2 441 SD 5 4.7 SD 0.4
85 327 SD 4 433 SD 16 4.8 SD 1.1
100 330 SD 2 443 SD 32 4.8 SD 2.1
b a
d c
Fig. 12. SEM micrographs of the fracture surface of a tensile specimen: (a) Al side; (b) MMC side; (c) high magnication of the at zones with the presence of ultrane
dimples in the Al alloy; (d) high magnication of the rough zones with the presence of ultrane dimples and reinforcement particles typical of the MMC.
0
50
100
150
200
250
300
350
400
450
1.00E+07 1.00E+06 1.00E+05 1.00E+04
M
a
x
i
m
u
m

S
t
r
e
s
s

[
M
P
a
]
Cycles [N]
10% failure probability
50% failure probability
90% failure probability
Experimental data
Fig. 13. Results of the axial fatigue tests at different failure probabilities on the dissimilar LFW joints.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 858
the MMC of the weld center and are characterized by the ductile
failure of the matrix with presence of ultra-ne dimples without
signicant decohesion or crack of the SiC particles. This morphology is
in agreement with the low degree of blending between the two
materials and suggests that fracture nucleates and develops in the
weld center both in the Al alloy and in the MMC.
3.4. Fatigue tests
The experimental fatigue data are reported in Fig. 13 with SN
curves at different failure probabilities. Fatigue strength at 10
7
cycles was equal to 190 MPa with a 50% probability of failure,
which is more than 40% of the ultimate strength of the AA2024
[2931]. This high fatigue resistance is due to the absence of
welding defects (Fig. 14), as also conrmed by the analysis of the
fracture surfaces.
SEM analysis of the fatigue fracture surfaces showed that
fracture nucleates and propagates in the TMAZ of the Al side,
and the nucleation sites mainly occurred on subsurface oxide
inclusions or intermetallic clusters which did not dissolve during
welding process (Fig. 15). The crack path is characterized by the
presence of classic fatigue striations but appears to be irregular,
possibly due to the grain morphology and orientation induced by
the LFW, which can deviate the crack front both macro- and
microscopically (Fig. 16) [33]. The nal overload failure of the
fatigue specimens often takes place along the welding line, with a
Fig. 14. SEM micrograph of the defect-free weld center.
b
a
d
c
Fig. 15. Fracture surface of a fatigue specimen: (a) low magnication SEM overview; (b) nucleation site of the crack; (c) back-scattered electron image of the nucleation
site showing the presence of a cluster of intermetallic compounds; (d) EDS spectrum of the intermetallic compounds.
Fig. 16. SEM micrographs of the crack path with irregular fatigue striations.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 859
morphology analogous to that already described for the tensile
specimen (Fig. 17).
4. Conclusions
Dissimilar joints between a 2024 Al alloy and a 2124/25 vol%
SiCp composite were successfully obtained by Linear Friction
Welding and their microstructural and mechanical characteriza-
tion was carried out.
Microstructural analyses of the joints showed substantially
defect-free joints. A substantial grain renement, due to the
concurrent effect of frictional heating and severe plastic deforma-
tion, was observed in the weld center, both on the Al and MMC
sides. A uniform particle distribution, whose size was unaffected
by the welding process, was observed in the MMC side. In the
TMAZ, the plastic ow led to signicant material brosity, with
grains elongated in the direction of expulsion of the material as
ash. The HAZ experienced some over-aging of the welded
materials.
The joint efciency, evaluated as the ratio between the joint
and the nominal AA2124 base material properties, was found to
be higher than 90% for both UTS and YS. The variation of the
welding parameters did not affect the tensile properties. Fatigue
strength at 10
7
cycles was found to be 190 MPa with a 50%
probability of failure. Fracture occurred in the MMC side in the
case of the tensile specimens and in the Al side in the fatigue
specimens.
Although further studies are needed, experimental results
demonstrated linear friction welding to be an attractive technol-
ogy for producing dissimilar joints between Al alloys and an
Al-based particle reinforced metal matrix composites (MMC). This
technique could contribute to widen the application range of this
class of composite, allowing for the localization in those compo-
nent parts where the MMCs specic properties would provide a
signicant performance improvement. The use of the LFW tech-
inque on this specic (or similar) combination of materials could
supply a way to produce components combining parts with high
damping capabilities with other parts characterized by high wear
resistance and greater stiffness, with applications, e.g., in the
aerospace or high-speed automatic machine sectors.
References
[1] O.T. Midling, O. Grong, Key Eng. Mater. 104107 (1995) 335372.
[2] A. Chidambaram, S.D. Bhole, Scr. Mater. 35 (3) (1996) 373378.
[3] M.D.B. Ellis, Int. Mater. Rev. 41 (2) (1996) 4158.
[4] R.S. Mishra, Z.Y. Ma, Mater. Sci. Eng. 50 (12) (2005) 178.
[5] R.P. Matrukanitz, ASM HandbookWelding, Brazing and Soldering 6 (1990)
528536.
[6] D. Bhattacharyya, M.E. Bowis, J.Y. Gregory, Proceedings of the ASM Interna-
tional Symposium on Machining of Composite Materials, Chicago, 1992,
pp. 4956.
[7] E.I. Kivineva, D. Olson, L. Matlock, Weld. J. 74 (3) (1995) 83s92s.
[8] K.L.B. Das, in: S.M. Lee (Ed.), International Encyclopedia of composites, vol. 2,
VCH, New York, 1990, pp. 460509.
[9] H.W. Vries, G. Ouden, Mater. Sci. Technol. 15 (2) (1999) 202206.
[10] A. Chidambaram, S.D. Bhole, Scr. Mater. 35 (3) (1996) 373378.
[11] G.J. Fernandez, L.E. Murr, Mater. Charact. 52 (2004) 6575.
[12] J.A. Wert, Scr. Mater. 49 (2003) 607612.
[13] I. Boromei, L. Ceschini, A. Morri, G. Garagnani, Metall. Sci. Technol. 24 (2006)
1221.
[14] L. Ceschini, I. Boromei, G. Minak, A. Morri, F. Tarterini, Compos. Part A 38
(2007) 12001210.
[15] L. Ceschini, I. Boromei, G. Minak, A. Morri, F. Tarterini, Compos. Sci. Technol.
67 (2007) 605615.
[16] G. Minak, L. Ceschini, I. Boromei, L. Ponte, Int. J. Fatigue 32 (1) (2010)
218226.
[17] F. Rotundo, L. Ceschini, A. Morri, T.S. Jun, A.M. Korsunsky, Compos. Part A 41
(2010) 10281037.
[18] A. Vairis, M. Frost, Mater. Sci. Eng. A271 (1999) 477484.
[19] P.P. Lean, L. Gilb, A. Urena, J. Mater. Process. Technol. 143144 (2003)
846850.
[20] J.M. Gomez de Salazar, M.I. Barrena, Mater. Sci. Eng. A 352 (2003) 162168.
[21] J.A. Wert, Scr. Mater. 49 (2003) 607612.
[22] ASTM E3-01. Standard Practice for Preparation of Metallographic Specimens.
ASM International, 2007.
[23] ISO/TTA 2. Tensile Tests for Discontinuously Reinforced Metal Matrix
Composites at Ambient Temperatures, 1997.
[24] A.H. Feng, B.L. Xiao, Z.Y. Ma, Compos. Sci. Technol. 68 (2008) 21412148.
[25] J. Guo, P. Gougeon, X.-G. Chen, Mater. Sci. Eng. A 553 (2012) 149156.
[26] J.M. Root, D.P. Field, T.W. Nelson, Metall. Mater. Trans. A 40 (2009)
21092114.
[27] X. Song, N. Baimpas, S. Harding, A.M. Korsunsky, Proceedings of the 4th
International Conference on Computational Methods for Coupled Problems in
Science and Engineering, Coupled Problems, 2011, pp. 13791387.
[28] S.C. Wang, M.J. Starink, N. Gao, Scr. Mater. 54 (2006) 287291.
[29] C. Genevois, D. Fabr egue, A. Deschamps, W.J. Poole, Mater. Sci. Eng. A 441
(2006) 3948.
[30] /http://www.matweb.comS.
[31] S.T. Amancio-Filho, S. Sheikhi, J.F. Dos Santos, C. Bolfarini, J. Mater. Process.
Technol. 206 (2008) 132142.
[32] U. Fuchs, K. Zimmermann, H.W. Sauer, K.H. Trautman, G. Biallas, Matwiss.
Werkstofftech. 39 (8) (2008) 531544.
[33] N. Kamp, N. Gao, M.J. Starink, I. Sinclair, Int. J. Fatigue 29 (2007) 869878.
Fig. 17. SEM micrographs of the zone where the nal overload failure of the fatigue specimen takes place: (a) low and (b) high magnication. The fracture morphology is
analogous to that of the tensile specimen.
F. Rotundo et al. / Materials Science & Engineering A 559 (2013) 852860 860

You might also like