Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Materials Characterization 108 (2015) 102–114

Contents lists available at ScienceDirect

Materials Characterization

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

Friction stir processing of an aluminum-magnesium alloy with


pre-placing elemental titanium powder: In-situ formation of an
Al3Ti-reinforced nanocomposite and materials characterization
F. Khodabakhshi a,⁎, A. Simchi b,c, A.H. Kokabi b, A.P. Gerlich d
a
Department of Materials Science and Engineering, School of Engineering, Shiraz University, Zand Boulevard, Shiraz, Iran
b
Department of Materials Science and Engineering, Sharif University of Technology, P.O. Box 11365-9466, Azadi Avenue, 14588 Tehran, Iran
c
Institute for Nanoscience and Nanotechnology, Sharif University of Technology, P.O. Box 11365-9466, Azadi Avenue, 14588 Tehran, Iran
d
Department of Mechanical and Mechatronics Engineering, University of Waterloo, Waterloo, ON, Canada

a r t i c l e i n f o a b s t r a c t

Article history: A fine-grained Al–Mg/Al3Ti nanocomposite was fabricated by friction stir processing (FSP) of an aluminum-magne-
Received 19 May 2015 sium (AA5052) alloy with pre-placed titanium powder in the stirred zone. Microstructural evolutions and formation
Received in revised form 13 August 2015 of intermetallic phases were analyzed by optical and electron microscopic techniques across the thickness section of
Accepted 27 August 2015
the processed sheets. The microstructure of the nanocomposite consisted of a fine-grained aluminum matrix
Available online 29 August 2015
(1.5 µm), un-reacted titanium particles (b40 µm) and reinforcement particles of Al3Ti (b100 nm) and Mg2Si
Keywords:
(b100 nm). Detailed microstructural analysis indicated solid-state interfacial reactions between the aluminum ma-
Friction stir processing trix and micro-sized titanium particles to form Al3Ti intermetallic phase. The hard inclusions were then fractured
Intermetallic nanocomposite and re-distributed in the metal matrix by the severe thermo-mechanical conditions imposed by FSP. Evaluation of
Aluminum–magnesium alloy mechanical properties by hardness measurement and uniaxial tensile test determined significant enhancement in
Al3Ti the mechanical strength (by 2.5 order of magnetite) with a high ductility (~22%). Based on a dislocation-based
Microstructure model analysis, it was suggested that the strength enhancement was governed by grain refinement and the pres-
Mechanical properties ence of hard inclusions (4 vol%) in the metal matrix. Fractographic studies also showed a ductile-brittle fracture
mode for the nanocomposite compared with fully ductile rupture of the annealed alloy as well as the FSPed speci-
men without pre-placing titanium particles.
© 2015 Elsevier Inc. All rights reserved.

1. Introduction matrix and the nanoparticles are achievable [12]. An alternative ap-
proach is friction stir processing (FSP), which has recently attracted a
Particulate-reinforced metal matrix composites (PMMCs) are of great attention for the fabrication of PMMCs at surfaces of base metals
great interest for structural and electrical applications considering [19–21]. Many feasibility studies have been performed on processing
their industrially feasible processing, low cost, and isotropic proper- of AMMNCs [19,21,22] as well as many other systems [13,23–31]. In
ties [1,2]. Advances in processing and applications of PMMCs relate to this process, a non-consumable tool with a concentric shoulder and spe-
utilizing nanometric reinforcements and refining the grain structure in cific pin design is plunged into the metal matrix workpiece [32]. By
order to activate higher strength with reasonable ductility [3,4]. Numer- adding second phase nanoparticles ahead of the tool during processing,
ous studies have been performed on the microstructure-property rela- surface modifications are attained [33,34]. Solid-state chemical reac-
tionship in various metal-matrix nanocomposites [1,2,5,6]. In tions may also occur upon processing, which can lead to the in-situ for-
particular, aluminum matrix nanocomposites (AMMNCs) have been in- mation of reinforcing particles [12]. The advantage of this process a
tensively studied [7,8]. These nanocomposites are fabricated by various more homogeneous microstructure with stronger interfacial bonding
techniques that can be classified in solid-state, liquid-state and deposi- of the nanoparticles to the metal matrix [8]. The role of FSP is to provide
tion processes [9–12]. One of the most widespread methods to prepare the following functions: (i) severe plastic deformation to promote
AMMNCs is the powder metallurgy route [13–18]. By employing this mixing and refining of the constituent phases in the material, (ii) elevat-
method, a fine distribution of the reinforcement nanoparticles in the ed temperature to facilitate the chemical reactions, and (iii) hot consol-
aluminum matrix with no or limited reactions between the metal idation to form a fully dense solid [35].
It is well known that aluminide intermetallics such as Al3Ni, Al3Ti,
Al3Fe, and Al2Cu have high wear resistance, hardness, specific strength,
⁎ Corresponding author. specific modulus, and excellent stability both at ambient and elevated
E-mail address: farzadkhodabakhshi83@gmail.com (F. Khodabakhshi). temperature [36]. Therefore, aluminum alloys reinforced with

http://dx.doi.org/10.1016/j.matchar.2015.08.016
1044-5803/© 2015 Elsevier Inc. All rights reserved.

Downloaded from http://www.elearnica.ir


F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 103

Fig. 1. (a) SEM image and (b) atomic absorption analysis result for the titanium particles. (c) Schematic representation of sample extraction from FSPed Al–Mg alloy and Al–Mg–Ti
composite for different microstructural analysis techniques, hardness measurements, and tensile testing.

aluminide particles have attracted much attention during the last two Recently, Hsu et al. [42,47] and Zhang et al. [45,46,54] have examined
decades due to their interesting mechanical properties [37–40]. Prepa- the in-situ formation of Al3Ti particles during FSP of Al/Ti powder mix-
ration of these nanocomposites by FSP is also attractive because the tures. More recently, we have also shown the in-situ formation of Al3Ti
aluminide reinforcing particles can efficiently be fractured, refined, nanoparticles upon FSP of an Al–Mg alloy with pre-placed TiO2 nano-
and dispersed by the thermo-mechanical processing conditions. Re- particles [55–62]. The role of the in-situ formed particles on the micro-
cently, Al–Al2Cu [41,42], Al–Al3Ni [29,43], Al–Al3Fe [44] and Al–Al3Ti structural evolutions and mechanical properties have been reported in
[42,45–47] composites have been successfully fabricated by employing details.
FSP on Al–Cu, Al–Ni, Al–Fe and Al–Ti elemental powder mixtures, re- In the present work, elemental titanium powder is employed instead
spectively. As compared to the other aluminide phases, aluminum- of TiO2 nanoparticles. Our previous results have shown that several
rich titanium intermetallic (Al3Ti) is very attractive due to its unique chemical reactions between the constituents (Al, Mg, Si, Ti and
properties, such as high melting point (~1350 °C), relatively low density O) occur during FSP that form different phases including MgO and
(~3.4 g.cm−3), high Young's modulus (~216 GPa), and low coarsening Al3Ti. Although enhanced mechanical strength was attained, magne-
rate at elevated temperatures [48]. Studies on Al–Al3Ti composites fab- sium oxidation reduces the age hardening potential of the aluminum
ricated by ingot metallurgy [49], mechanical alloying [50,51], rapid so- matrix. Analysis of the strengthening mechanisms were also difficult
lidification [52] and combustion synthesis [53] have shown the as multiple and complex phases were formed. The aim of this work is
potential of Al3Ti particles to improve the stiffness of aluminum alloys. to study the fragmentation and dissolution of titanium particles in the

Table 1
The atomic absorption chemical analysis results from the initial titanium reinforcing particles (conditions for atomic absorption analysis were set as; accelerating voltage of 20 kV, beam
current of 500 pA, magnification of 60, live time of 40 s, preset time of 40 s, Nb channels of 2048, Ev/channel of 10, offset of 0 keV, and width of 20 keV).

Element Line Int Error K Kr W% A% ZAF Formula Ox% Pk/Bg Class LConf HConf Cat#

Ti Ka 1952.7 2.7983 1.0000 1.0000 100.00 100.00 1.0000 0.00 125.24 A 99.28 100.72 0.00
1.0000 1.0000 100.00 100.00 0.00 0.00

# These numbers are the outputs of atomic absorption chemical analysis.


104 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

compared with the base metal, as well as the nanocomposites prepared


by pre-placed TiO2 nanoparticles in previous studies.

2. Experimental procedure

2.1. Materials

Commercial aluminum–magnesium alloy (5052-H32 rolled and


tempered sheet) with a chemical composition of Al–2.23 Mg–
0.292Fe–0.163Cr–0.147Si–0.114Mn–0.0077Zn–0.0065Cu (in wt.%) and
thickness of 5 mm was supplied from Arak Aluminum Company, Arak,
Iran. The sheet was cut into 210 × 70 mm2 specimens and solution-
treated by annealing at 500 °C for 2 h followed by water quenching. Ti-
tanium powder with a commercial purity of 99.9%, and an average par-
ticle diameter of 40 μm was purchased from Merck Company
(Germany). Morphological and chemical analyses of the titanium pow-
der are presented in Figs. 1a,b and Table 1, respectively.

2.2. Friction stir processing

A groove with a width of 1.2 mm and depth of 3.5 mm was machined


in the middle of the sheets and filled with the titanium powder (~0.6 g).
After tapping with hand, a capping FSP pass was performed by employing
Fig. 2. Stereographic macro-images display the stir zone flow pattern of friction stir
processed (a) Al–Mg alloy and (b) Al–Mg–Ti composite. a tool with a shoulder (12 mm diameter) having no pin. A modified mill-
ing machine was used to conduct the FSP experiments. By trail-and-error,
it was found that the suitable parameters for the accomplishment of the
Al–Mg matrix during severe plastic deformation and heat flow upon capping pass without severe vibrations were: rotational speed (w) of
FSP. The effect of magnesium on the in-situ chemical reactions and its 1125 rpm and traverse velocity (v) of 30 mm/min. The FSP tool was
role on the formation of intermetallic phases are elaborated. Additional- made of H13 steel with a concave shape shoulder (diameter of 18 mm
ly, the mechanical properties of the nanocomposites are examined and and concave angle of 2.5°) and a cylindrical threaded pin (height of

Fig. 3. Optical microstructures for BM, HAZ, TMAZ and SZ of FSPed Al–Mg alloy at different magnifications.
F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 105

Fig. 4. Optical microstructures for BM, HAZ, TMAZ and SZ of FSPed Al–Mg–Ti composite at different magnifications.

4 mm, diameter of 5 mm, M5 threads). A counterclockwise rotational kept around 2.5°. Multi-passes FSP was performed to incorporate the
speed of 1200 rpm and a traverse velocity of 100 mm/min were used. Ti particles into the Al matrix uniformly and to pursue Al/Ti solid-state
The nuting angle with respect to the normal axis of sample plane was chemical reactions. The processing conditions were not changed for

Fig. 5. Optical micrographs showing the distribution of reinforcement particles within the grain interior as well as along the grain boundaries of the Al–Mg metal matrix.
106 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

Fig. 6. FE-SEM images display the titanium particles dispersion through the aluminum matrix.

repeated passes but 100% overlap was considered. The pressing depth concentration profiles for the dispersed initial reinforcing particles
of tool shoulder for each FSP passes was set to ~0.2 mm via a displace- and newly formed intermetallic phases were investigated by energy
ment control unit. Based on the groove dimensions and the weight of dispersive X-ray spectroscopy (EDS) analysis. For characterization
the Ti powder, the amount of the Ti incorporated in the matrix was and identifying in situ reactions and new formed phases scanning
about ~ 3 vol.%. The base Al–Mg alloy was also FSPed at the same pro- transmission electron microscopy (STEM; JEM-2100 F STEM, JEOL,
cessing conditions to explore the role of Ti addition. The results of mi- Japan) and high resolution-transmission electron microscopes
crostructural evaluations and mechanical examinations were also (HRTEM; Philips, Netherlands) were utilized. The sample prepara-
compared with the nanocomposites prepared by pre-placing TiO2 tion approach for FESEM samples was similar to that used for STEM
nanoparticles [55,58–60,62] in order to highlight the effect of additive and HRTEM samples comprehensively explained in prior works [58,
type. 62].

2.3. Microstructural characterization 2.4. Mechanical testing

The processed samples were cross-sectioned perpendicular to the Details of samples extraction for different mechanical testing are in-
FSP direction by electron discharge machining to examine the micro- dicated in Fig. 1c, schematically. The hardness profile along different
structure as schematically shown in Fig. 1c. Standard metallographic zones of the FSPed specimens was recorded on the metallographic
procedures were employed through grinding with SiC papers followed cross-sections utilizing a Vickers micro-hardness indenter (Bohler,
by polishing with diamond pasts. The polished cross-sections were Germany). The measurement was performed underneath the top sur-
chemically etched by modified Poulton's reagent consisting of two solu- faced with 2 mm distance. The applied load was 200 g with a 15 s
tions (H2O (1 ml)–HNO3 (6 ml)-HF (1 ml)–HCl (12 ml)/HNO3 (25 ml)– dwell time. Thin rectangular specimens (3 mm thickness) with a gage
H2CrO4 (1 g)–H2O (10 ml)). Microstructural studies of different zones length of 32 mm and a width of 6 mm were prepared in longitudinal
including base metal (BM), heat affected zone (HAZ), thermo- and perpendicular directions (relative to the FSP line) according to
mechanical affected zone (TMAZ) and stir zone (SZ) were examined ASTM E8M standard. The stir zone was located in the gage length. A
using stereographic (Nikon, USA) and optical (Olympus, Germany) mi- Universal Tensile Loading Machine (Model H10K, Tinius Olsen, USA)
croscopes. The average grain size of different regions was measured with an initial strain rate of 5 × 10−4 s−1 was employed. Triplicate spec-
by the mean linear intercept method. Further microstructural exam- imens were examined for each conditions and materials, and the aver-
inations were performed by field emission-scanning electron age values were reported. After fracturing, the cross-section was
microscopy (FE-SEM, JEOL 7600, Japan). Line-scan and elemental observed by FE-SEM.
F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 107

3.2. Analysis of secondary phases

Optical micrographs and FE-SEM images in Figs. 5 and 6 respectively,


show details of microstructural features in SZ for the FSPed Al–Mg alloy
in the presence of Ti particles. Fig. 7 shows the high magnification FE-
SEM images of the FSPed composite. Relatively coarse Ti particles
(white contrast particles in the FE-SEM images and brown contrast in
the optical micrographs) and round Al3Ti particles with a diameter
b100 nm were observed. The volume fraction of the Ti particles was ap-
proximately 2% based on quantitative microstructural analysis of FE-
SEM images in Figs. 6 and 7. Since, Al3Ti is the only reaction product be-
tween Al and Ti elements, the amount of in situ formed Al3Ti intermetal-
lic nanoparticles was calculated from the volume fraction of residual Ti
as 4 vol.% based on a mass balance calculation. Also, it should be noted
that the Al3Ti intermetallic phases were observed close to the Ti parti-
cle/matrix interfaces (Fig. 7a,b), supporting the occurrence of solid-
state chemical reactions between the aluminum matrix and titanium
particles. Fig. 8 shows FE-SEM images and the corresponding elemental
maps for the processed specimen. For this studied composite system, as
seen in Fig. 8a at low magnification, no Mg-rich particles were found,
and Mg was distributed into the Al matrix, however, the distribution
of Mg in the matrix was nonhomogeneous. Although in elemental
maps of Fig. 8b at high magnification, some weak intensities of ultra-
fine Mg rich particles were detected. Based on the previous research
on the Al–Mg–Ti complex system [63,64], these particles maybe FCC
structured Ti2Mg3Al18, however in the present work these are not Al
and Ti rich. Therefore, these nano-metric round phases are likely
Mg2Si particles which formed as a result of re-precipitation of some sol-
ute Mg atoms during FSP, since there is 0.147% Si in the base metal and
this would be comparable to the AA6000 series alloy which are mainly
strengthened by Mg2Si [18,65]. These nano-scaled Mg2Si precipitates
can contribute to secondary strengthening of the Al matrix around the
Al3Ti nanoparticles. Line-scan EDS analysis results from different size
particles combined with FE-SEM images in Fig. 9 are in agreement
with the obtained findings from elemental mappings analysis. Micro-
structural features including the sub-grains structure, dislocations ori-
Fig. 7. A magnified view by FE-SEM images at large magnifications demonstrate the Al–Mg entation, and twinning produced by severe plastic deformation of Al
matrix/Ti particles interface and formation of in situ Al3Ti nanoparticles. The white phase matrix at SZ are shown in STEM and HR-TEM images of Fig. 10a–c.
is titanium, the light gray phase is Al3Ti, and the dark matrix is Al–Mg alloy.
High magnification STEM images from the in situ formed Al3Ti nanopar-
ticles in the Al matrix are displayed in Fig. 10d–f. An HR-TEM image
3. Results from the Al matrix/Al3Ti interface shown in Fig. 10g, indicating a good
crystallographic matching between the newly formed reinforcement
3.1. Materials flow and grain structure phase during processing and the metal matrix. Corresponding selected
area diffraction (SAD) spot patterns for the in situ Al3Ti phase and the
The patterns of material flow for the processed Al–Mg alloy without severely deformed Al matrix from [110] zone axis are shown in
and in the presence of Ti particles are shown in Fig. 2a and b, respective- Figs. 10h,i, respectively. By indexing the SAD pattern of Fig. 10h, it was
ly. It is apparent that the processed zone of both samples exhibited an confirmed that those nano-metric Ti reach particles detected by FE-
onion ring flow structure, while the intensity and sharpness of the SEM analysis are Al3Ti phase. Typical reported results of microstructural
banded structure is increased significantly when the Ti powder is analysis for the processed sample with TiO2 nanoparticles in the previ-
added. It is clear that the incorporation of Ti particles during FSP has dra- ous studies [58–60,62] were revealed the in situ formation of MgO
matically influenced the material flow and geometry of the stir zone. It and Al3Ti phases, as the kinetics for Mg2Si precipitates formation were
seems that a homogenous distribution from Ti micro-particles through restricted, attributed to a higher chemical affinity of magnesium to oxy-
SZ was achieved after employing of FSP for four passes. Figs. 3 and 4 gen than that of titanium.
show microstructural features of different zones (BM, HAZ, TMAZ, and
SZ) formed upon FSP of Al–Mg alloy without and with Ti particles, re- 3.3. Mechanical properties
spectively. The base metal consisted of a relatively coarse grain struc-
ture (~ 49 μm), while a relatively uniform microstructure with an Micro-hardness profiles across SZ for the Al–Mg alloy and Al–Mg–Ti
average size of ~ 20 μm were formed after FSP (Fig. 3a–c), indicating system are plotted in Fig. 11a. The mean Vickers hardness values for dif-
that dynamic recrystallization (DRX) occurred during FSP. Microstruc- ferent zones of BM, HAZ, TMAZ, and SZ are reported in Table 2. The base
tural studies revealed that the addition of secondary particles led to a metal exhibited a uniform hardness across the cross section with an av-
significant grain refinement. The average size of grains for the processed erage value of ~51 HV. An increased (~16%) in the hardness of SZ was
alloy in the presence of Ti particles was about 1.5 μm, as demonstrated attained after FSP without the addition of the secondary particles. How-
in Figs. 4a–c and 5a–d. It should be noted that, the FE-SEM results in ever, the hardness was ramp up in SZ to ~131 HV (~156% increment)
Figs. 6–9 and STEM images of Fig. 10 should be considered supplemen- when the titanium particles were introduced. The heterogeneous hard-
tary microscopy in determining the mean size of SZ high angle grain ness profile was attributed to the presence of distributed coarse Ti par-
structure in the composite sample. ticles (Figs. 5 and 6). Fig. 11b shows the engineering stress-strain curves
108 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

Fig. 8. FE-SEM images combined with the elemental mapping analysis results.

Fig. 9. Line scan analysis from reinforcement particles with three different sizes.
F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 109

Fig. 10. (a) Bright-, (b) dark-field STEM, and (c) HR-TEM images from the hot severely deformed structure of SZ showing dislocations, twin boundaries, sub-grains, and the other
microstructural features. (d–f) High magnification STEM images from the in situ formed Al3Ti nanoparticles, and (g) HR-TEM image from the Al–Mg matrix/Al3Ti interface. The SAD
spot patterns for (h) the in situ Al3Ti phase and (i) the deformed Al matrix.

for the examined materials. The main mechanical characteristics obtain- remarkable. Much lower mechanical properties were measured upon
ed from these graphs are summarized in Table 2. The annealed Al–Mg transverse tensile testing as the specimens were fractured from the
alloy exhibited relatively low yield stress (YS, ~ 68 MPa) and ultimate SZ/TMAZ interface.
tensile strength (UTS, ~ 180 MPa) with high elongation (~ 29.5%). A
slight enhancement in the strength and elongation was achieved after 3.4. Fracture behavior
FSP of the Al–Mg alloy due to grain structural refinement because of op-
erating DRX mechanism occurred in the SZ during the processing. The The fracture surface of the tensile tested specimens was observed by
addition of Ti particles enhanced the longitudinal tensile strength to FE-SEM to study the failure mechanisms. Fig. 12 shows typical micro-
values of 151 MPa (yield) and 246 MPa (ultimate) with a decrease in graphs of the fracture surfaces. The annealed Al–Mg alloy exhibited a
the tensile elongation to 22%, which are more significant are much im- ductile rupture behavior with characteristics of gray fibrous surface
provements than the measured values in previous works [58–60,62] with dimples and the presence of shear zones (Fig. 12a–c). As shown
for the FSPed Al–Mg–TiO2 system. By incorporating the ex situ Ti in the related stress–strain graph of Fig. 11b, significant deformation
micro-particles and subsequently forming new in situ Al3Ti nano- and considerable amount of necking was observed in this sample. No
particles during FSP process, dislocation density around the reinforcing major difference was observed for the FSPed Al–Mg alloy except finer
particles is increased significantly while simultaneously refining the dimple sizes (Fig. 12d–f), which could be related to its finder grain
grains due to elastic mismatch and particle stimulated nucleation mech- structure [60]. The presence of a gray fibrous surface with dimples in
anisms, respectively [46,47,54,66]. As a result, the hardness and tensile these non-reinforced samples demonstrates the fracture mode was a
strength of the processed nanocomposite was enhanced and the elon- shear ductile rupture. Deep and elongated dimples are the results of nu-
gation reduced as compared to the initial and processed Al–Mg alloys. cleation of micro voids, their growth and finally coalescence of dimples
However, it can be noted that the magnitude of elongation is which are affected by shear stress [67]. It was found that the presence of
110 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

refined to equiaxed grains with size of ~20.3 μm after employing FSP for
four passes without introducing the reinforcement particles. It is sug-
gested that once a given volume of plastically deformed material was
located the outside the deformation zone of the rotating tool pin, it
has coarsened very quickly under the influence of heat [69]. For both
processed samples, a sharp transition in grain size can be noted moving
from the BM toward the SZ grain structure, with drastic refinement
clearly visible. In the Al–Mg alloy with an initial solution-treated condi-
tion, discontinuous dynamic recrystallization (DDRX) can occur during
the hot deformation conditions involved during FSP, and the tendency
for DDRX to occur can be attribute to the moderate to low stacking
fault energy in this alloy [58,60]. In DDRX, new, dislocation-free grains
form at sites such as prior grain boundaries, deformation band inter-
faces or boundaries of newly recrystallized grains when a critical dislo-
cation density is achieved within the grains [70]. The driving force for
initiation of recrystallization is the strain-induced build up of stored
strain energy in the form of fine dislocation cells and a high density of
mobile dislocations [71]. After FSP of the Al–Mg–Ti composite system,
as a result of the in situ reactions promoting intermetallic phase forma-
tion, the average grain size of SZ was reduced significantly down to
~1.5 μm, as shown in Figs. 5 and 8a. Insoluble nanoparticles can provide
suitable sites to enhance the nucleation rate of new grains at the begin-
ning of DDRX by particulate stimulated nucleation (PSN) [72], and sub-
sequently the rate of grain boundary migration would have been
hindered by Zener–Holloman pinning (ZHP) mechanism also contribut-
ed by the fractured Ti powder particles and other fine nucleated parti-
cles which will be discussed shortly [73]. Both of these PSZ and ZHP
mechanisms operate in concert and lead to enhanced grain refinement
during FSP. Hence, multiple microstructural and phase transformational
phenomena are operating during FSP of Al–Mg–Ti composite, which led
to a grain structure in the SZ which seems to be bimodal in size
distribution.
Fig. 11. (a) Micro-hardness profiles across thickness section of annealed Al–Mg alloy be-
fore and after FSP, and FSPed Al–Mg–Ti composite. (b) Engineering stress-strain curves 4.2. Formation mechanisms of intermetallic phases
display transverse and longitudinal tensile flow behaviors of FSPed Al–Mg–Ti composite
in comparison to the annealed Al–Mg alloy before and after FSP. As revealed in previous works for FSPed Al–Mg–TiO2 system [59,62],
the in situ solid-state reactions during FSP is controlled by diffusion of
reinforcement particles have a significant influence on the fracture sur- Mg in the Al matrix/TiO2 nanoparticles interface and as a result, the
face features. As shown in Fig. 12g–i, the fracture surface of the Al–Mg MgO phase is formed according to deformation-assisted interfacial
alloy with the addition of Ti particles showed a ductile–brittle failure chemical reaction mechanism. As the reaction proceeds, titanium
mechanism. Fine dimples without significant shear tearing were ob- atoms will dissolute in the aluminum matrix to form Al3Ti at the parti-
served. Some cracked particles were found on the fracture surface of cles–matrix interface by deformation-assisted solution–precipitation
composite sample (gray arrows in Fig. 12g–i). Particles were seen in mechanism. However, different mechanisms of in situ phase formation
the core of most dimples, supporting the role of particle-matrix inter- are observed in this study for the FSPed Al–Mg–Ti intermetallic system.
face on the rupturing [68]. EDS analysis determined that those particles High magnification FE-SEM images in Fig. 7 provide some indication of
were mostly Al3Ti and un-reacted Ti particles, as the results are present- the mechanism for reactive diffusion between Al and Ti. At first, Al3Ti
ed in Fig. 13. The related regions for these EDS analysis are indicated on phase formed at the outer layer of the initial Ti micro-particles and a
the FE-SEM fractographs of different samples in Fig. 12c,f,i. structure like core-shell produced. The difference in the diffusion coeffi-
cients between Ti in Al during the Al–Ti inter-diffusion would create va-
4. Discussion cancies in the Al as a result of Kirkendall's effect, thereby causing a
volume expansion in Ti [47,66]. Furthermore, the density of Al3Ti is
4.1. Effect of secondary particles on the grain structure lower than that of Ti, hence the formation of Al3Ti would cause a consid-
erable volume expansion. The volume expansion would produce tensile
As shown in Fig. 3, the initial annealed Al–Mg base metal posses coarse stresses in the outer layer of the particles and compressive stress in the
grain structure with an average size ~49 μm, and this microstructure was core of the particles, and this resulted in the initiation of some micro-

Table 2
Tensile and indentation properties for different zones (BM, HAZ, TMAZ, and SZ) of the friction stir processed Al–Mg alloy and Al–Mg–Ti composite: Elastic modulus (E, MPa); Yield stress
(YS, MPa); Ultimate tensile strength (UTS, MPa); Elongation (e, %); Fracture stress (FS, MPa); Micro-hardness (HV, Vickers).

Properties Tensile properties Vickers hardness

Sample Condition Orientation E YS UTS e FS BM HAZ TMAZ SZ

Al–Mg Annealed Longitudinal 70 68 180 29.5 159 51.1 ± 0.6 – – –


Al–Mg FSP Longitudinal 71 65.3 194.4 33.2 158.8 51.1 ± 0.6 51.1 ± 0.6 51.6 ± 0.8 59.5 ± 2.0
Al–Mg–Ti FSP Transverse 72 92.5 195.0 14.0 183.4 51.1 ± 0.6 51.1 ± 0.6 57.8 ± 8.7 130.6 ± 20.6
Al–Mg–Ti FSP Longitudinal 76 151.5 246.2 22.2 209 – – – –
F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 111

Fig. 12. FE-SEM images show the fracture surfaces of (a–c) initial Al–Mg alloy, (d–f) FSPed Al–Mg alloy, and (g–i) FSPed Al–Mg–Ti composite at (a, d, g) 200×, (b, e, h) 1000×, (c, f, i) and
2000× magnifications.

cracks in the Al3Ti as the reaction proceeded during multi-pass FSP where Ec, Em, and Ep are the elastic modulus of the composite, metal ma-
(STEM images of Fig. 10d–f). The number of micro-cracks would be in- trix, and reinforcing particles, respectively, η is an adjustable parameter,
crease with the volume expansion increases. The large brittle Al3Ti and V is the volume fraction of the particles. For the prepared Al–Mg–
blocks can be easily broken into smaller particles with sizes b100 nm Al3Ti nanocomposite, Em = 70 GPa and Ep = 216 GPa [75]. By taking
due to the existence of these micro-cracks and imposed intense plastic η = 1, the calculated Young's modulus of FSPed in situ intermetallic
deformation during continuation of FSP process. According to the result nanocomposite with 4% volume fraction of Al3Ti nanoparticles is report-
of elemental maps (Fig. 8) and EDS analysis (Fig. 9), no Mg-containing ed in Table. 3. As seen, the Measured Young's modulus of Al–Mg–Al3Ti
intermetallics were formed in the FSPed Al–Mg–Ti sample and Mg dis- nanocomposite agrees well with the prediction of the Halpin–Tsai equa-
solved into the Al matrix, although its distribution was non-uniform tion. This suggests that there is good bonding strength between Al3Ti
due to formation of nano-metric Mg2Si precipitates. The presence of sol- nanoparticles and the Al matrix so that in situ formed Al3Ti nanoparti-
ute Mg would facilitate the Al–Ti inter-diffusion at the Al–Ti interface to cles can contribute effectively to load sharing in the nanocomposite.
form Al3Ti particles as a result of the Al–Ti reaction. The severe plastic
deformation during FSP, would promote a thermo-mechanically acti- 4.3.2. Yield strength and strengthening mechanisms
vated condition which removed the Al3Ti particles from the Al–Ti inter- Four possible strengthening micro-mechanisms which may operate
face and distributed them into the Al–Mg metal matrix uniformly. in particulate-reinforced metal matrix nanocomposites are [76]; (i)
grain and substructure strengthening, (ii) Orowan strengthening, (iii)
4.3. Modeling of mechanical properties quench hardening resulting from the dislocations generated to accom-
modate the differential thermal contraction between the reinforcing
4.3.1. Elastic modulus nanoparticles and the metal matrix, and (iv) work hardening due to
With the incorporation of titanium particles through Al–Mg metal the strain misfit or elastic modulus mismatch between the elastic rein-
matrix during FSP, the Young's modulus of prepared in situ intermetallic forcing particles and the plastic matrix. Due to the low aspect ratio of re-
reinforced nanocomposite was increased noticeably in comparison to inforcing phases in particulate nanocomposites, the contribution of load
Al–Mg base metal or FSPed Al–Mg alloy without pre-placed particles transferring mechanism is negligible. By considering the superimposed
as shown in Table 2, which can be attributed to the high elastic modulus effects of Hall–Petch strengthening (σH − P), Orowan strengthening
of Al3Ti phase. The Young's modulus of particulate-reinforced compos- (σOR), coefficient of thermal expansion (σCTE) and elastic modulus mis-
ites can be predicted from the Halpin–Tsai equation as follows [74]: match (σEM), the strength of the matrix can be expressed as [77]:

σ c ¼ σ m þ σ H−P þ σ OR þ σ CTE þ σ EM ð2Þ


 
Em Ep ð1 þ ηV Þ þ ηEm ð1−V Þ where σm is the strength of the unreinforced metal matrix (~50 MPa for
Ec ¼ ð1Þ
Ep ð1−V Þ þ Em ðη þ V Þ the examined Al–Mg alloy). The grain size has a strong influence on
112 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

where D is the grain size, k is the strengthening proportional constant


(~ 70 MPa.μm−0.5), M is the Taylor factor (~ 3 for face-centered cubic
polycrystals), τOR is the shear strength associated with reinforcing parti-
cles, β is a constant, G is the matrix shear modulus (~26.2 GPa), b is the
Burgers vector (~0.286 nm), ρCTE and ρEM are geometrical necessary dis-
locations density due to CTE and EM respectively, ν is Poisson's ration
(~0.345), dp is the diameter of Al3Ti nanoparticles (~100 nm), r0 is the
dislocation core radius (=4 b), A is a geometric constant, Δα is the dif-
ference in CTE, ΔT is the difference between test and processing temper-
atures, and ε is the strain in EM [78]. To estimate the contribution of
Orowan strengthening by Al3Ti nanoparticles, it is assumed that these
particles are spherical and uniformly distributed. The interparticle spac-
ing, λ, of Al3Ti nanoparticles with diameter of dp and volume fraction of
Vp can be calculated from the following equation [79].

" sffiffiffiffiffiffi! # rffiffiffi!


1 π 2
λ¼ −2 dp ð5Þ
2 Vp 3

Based on the microstructural details explained before, the contribu-


tions of different strengthening mechanisms for FSPed Al–Mg alloy and
Al–Mg–Al3Ti nanocomposite were calculated and presented in Table 3.
Moreover, the total yield strengths of these processed materials were
predicted and compared by the estimated values from experimental
measurements. As shown, the high strength of in situ intermetallic
nanocomposite prepared by FSP can be mainly attributed to the grain
size strengthening which results from fine grain size of Al–Mg metal
matrix, and the Orowan strengthening which is due to high volume
fraction of nano-metric sized Al3Ti particles. As explained in the men-
tioned modeling procedure floc-shaped particles were ignored and as-
sumed that all of nano-particles homogeneously distributed, but as
Fig. 13. EDS chemical analysis correspond to the fracture surfaces of (a) initial Al–Mg alloy, revealed by microstructural studies the fraction of floc-shaped particles
(b) FSPed Al–Mg alloy, and (c) FSPed Al–Mg–Ti composite. These analyses's performed in the fabricated nanocomposites is considerable. Therefore, some dif-
from the indicated regions in FE-SEM fractographs of Fig. 12c, f, and i, respectively.
ferences between expected strength through modeling and experimen-
tally measured strength as presented in Table 3 can be reasonable.
metal strength since the grain boundaries can hinder the dislocation
movement. This is due to the different orientation of adjacent grains 5. Conclusions
and to the high lattice disorder characteristic of these regions, which
prevent the dislocations from moving in a continuous slip plane [76]. An Al–Mg/Al3Ti (b100 nm, 4 vol%) nanocomposite was fabricated by
The so-called Orowan mechanism consists in the interaction of nano- friction stir processing of an Al–Mg alloy (AA5052) with pre-placed Ti
particles with dislocations. The non-shearable ceramic reinforcement particles. The formation mechanism of intermetallic particles during
particles pin the crossing dislocations and promote dislocations bowing FSP, microstructural evolutions and mechanical properties of the nano-
around the particles (Orowan loops) under external load [77]. The mis- composites were studied and compared with the un-reinforced alumi-
match in coefficient of thermal expansion (CTE) and in elastic modulus num alloy. Microstructural studies determined that the average grain
(EM) between the reinforcements and the metal matrix is accommo- size of the Al–Mg alloy decreased from ~49 µm down to 20 µm by
dated during material cooling and straining by the formation of geomet- employing FSP. Pre-placing of titanium particles accelerated the grain
rically necessary dislocations [77]. Therefore, the final strength of the refinement, so that a fine-grained metal matrix was attained. The aver-
composite can be evaluated by summing the above contributions relat- age grain size of the composite was about 1.5 µm but the size distribu-
ed to the single strengthening effects as indicated below [76]: tion was broad with a sign of bimodal distribution. The large plastic
strain and high temperature introduced by FSP also facilitated solid-
pffiffiffi qffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffi
k state chemical reactions between Al and Ti to form Al3Ti particles. The
σ c ¼ σ 0 þ pffiffiffiffi þ Mτ OR þ 3βGb ρCTE þ ρEM ð3Þ
D size of the hard inclusions was smaller than 100 nm with an approxi-
mate volume fraction of 4%. Ultra-fine Mg2Si particles were also detect-
rffiffiffi ! (sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffi) ed in the microstructure. Due to the presence of hard inclusions, the
k 0:81GbM 2 dp pffiffiffi AΔαΔTV p 6V p
σ c ¼ σ 0 þ pffiffiffiffi þ ln þ 3βGb  þ 3
ε mechanical strength of the nanocomposite was considerably increased
D 2πð1−νÞ1=2 λ 3 r0 bdp 1−V p πdp
compared to the unreinforced alloy with the material still exhibited rea-
ð4Þ sonably high tensile elongation (~22%). Model analysis revealed the

Table 3
Predicted Young's modulus and contribution of different strengthening mechanisms to the yield strength of friction stir processed Al–Mg alloy and Al–Mg–Ti intermetallic composite sys-
tem: Ec,p, Predicted Young's modulus; Ec,m, Measured Young's modulus; σyc,p, Predicted tensile yield strength; σyc,m, Measured tensile yield strength.

Properties Elastic modulus Yield strength

Sample Ec,p(GPa) Ec,m(GPa) σ0(MPa) σH − P(MPa) σOR(MPa) σCTE(MPa) σEM(MPa) σyc,p(MPa) σyc,m(MPa)

FSPed Al–Mg alloy 70 71 50 15.5 – – – 65.5 65.3


FSPed Al–Mg–Ti composite 75 76 50 57.2 54.6 30.4 5.2 197.4 151.5
F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114 113

importance of grain refinement and distributing nano-metric hard in- [27] M.A. Moghaddas, S.F. Kashani-Bozorg, Effects of thermal conditions on microstruc-
ture in nanocomposite of Al/Si3N4 produced by friction stir processing, Mater. Sci.
clusions on the mechanical properties of Al–Mg/Al3Ti (b100 nm, Eng. A 559 (2013) 187–193.
4 vol%) nanocomposite. Fractographic studies also showed a combined [28] A. Mostafapour Asl, S.T. Khandani, Role of hybrid ratio in microstructural, mechani-
ductile–brittle fracture mode for the nanocomposite compared to the cal and sliding wear properties of the Al5083/Graphitep/Al2O3p a surface hybrid
nanocomposite fabricated via friction stir processing method, Mater. Sci. Eng. A
fully ductile rupture of the annealed Al–Mg alloy as well as the FSPed 559 (2013) 549–557.
Al–Mg alloy without pre-placing titanium particles. [29] J. Qian, J. Li, J. Xiong, F. Zhang, X. Lin, In situ synthesizing Al3Ni for fabrication of
intermetallic-reinforced aluminum alloy composites by friction stir processing,
Mater. Sci. Eng. A 550 (2012) 279–285.
[30] C.M. Rejil, I. Dinaharan, S.J. Vijay, N. Murugan, Microstructure and sliding wear be-
Acknowledgments havior of AA6360/(TiC and B4C) hybrid surface composite layer synthesized by fric-
tion stir processing on aluminum substrate, Mater. Sci. Eng. A 552 (2012) 336–344.
[31] S. Sahraeinejad, H. Izadi, M. Haghshenas, A.P. Gerlich, Fabrication of metal matrix
Financial support from the Iranian National Elite's Foundation is
composites by friction stir processing with different particles and processing param-
greatly appreciated. eters, Mater. Sci. Eng. A 626 (2015) 505–513.
[32] F. Khodabakhshi, M. Haghshenas, S. Sahraeinejad, J. Chen, B. Shalchi, J. Li, A.P.
Gerlich, Microstructure-property characterization of a friction-stir welded joint be-
tween AA5059 aluminum alloy and high density polyethylene, Mater. Charact. 98
References (2014) 73–82.
[33] X. He, F. Gu, A. Ball, A review of numerical analysis of friction stir welding, Prog.
[1] C. Suryanarayana, Synthesis of nanocomposites by mechanical alloying, J. Alloys Mater. Sci. 65 (2014) 1–66.
Compd. 509 (2011) 229–234. [34] R. Nandan, T. DebRoy, H.K.D.H. Bhadeshia, Recent advances in friction-stir welding –
[2] C. Suryanarayana, N. Al-Aqeeli, Mechanically alloyed nanocomposites, Prog. Mater. process, weldment structure and properties, Prog. Mater. Sci. 53 (2008) 980–1023.
Sci. 58 (2013) 383–502. [35] A. Simar, Y. Bréchet, B. de Meester, A. Denquin, C. Gallais, T. Pardoen, Integrated
[3] J. Hemanth, Development and property evaluation of aluminum alloy reinforced modeling of friction stir welding of 6xxx series Al alloys: process, microstructure
with nano-ZrO2 metal matrix composites (NMMCs), Mater. Sci. Eng. A 507 (2009) and properties, Prog. Mater. Sci. 57 (2012) 95–183.
110–113. [36] S.K. Das, L.A. Davis, High performance aerospace alloys via rapid solidification pro-
[4] F. He, 6 – Ceramic Nanoparticles in Metal Matrix Composites, in: R. Banerjee, I. cessing, Mater. Sci. Eng. 98 (1988) 1–12.
Manna (Eds.), Ceramic Nanocomposites, Woodhead Publishing 2013, pp. 185–207. [37] J.M. Wu, S.L. Zheng, Z.Z. Li, Thermal stability and its effects on the mechanical prop-
[5] R. Bogue, Nanocomposites: a review of technology and applications, Assem. Autom. erties of rapidly solidified Al–Ti alloys, Mater. Sci. Eng. A 289 (2000) 246–254.
31 (2011) 106–112. [38] R. Lerf, D.G. Morris, Mechanical alloying of Al–Ti alloys, Mater. Sci. Eng. A 128 (1990)
[6] V. Provenzano, R.L. Holtz, Nanocomposites for high temperature applications, Mater. 119–127.
Sci. Eng. A 204 (1995) 125–134. [39] L. Zhang, B.I. Wu, Y.H. Zhao, X.H. Du, Exploration of Al-based matrix composites re-
[7] C. Suryanarayana, Mechanical alloying and milling, Prog. Mater. Sci. 46 (2001) inforced by hierarchically spherical agents, Int. J. Miner. Metall. Mater. 20 (2013)
1–184. 796–801.
[8] C. Suryanarayana, E. Ivanov, 3 – Mechanochemical Synthesis of Nanocrystalline [40] S.C. Tjong, Z.Y. Ma, Microstructural and mechanical characteristics of in situ metal
Metal Powders, in: I. Chang, Y. Zhao (Eds.), Advances in Powder Metallurgy, matrix composites, Mater. Sci. Eng. R 29 (2000) 49–113.
Woodhead Publishing 2013, pp. 42–68. [41] C.J. Hsu, P.W. Kao, N.J. Ho, Ultrafine-grained Al–Al2Cu composite produced in situ by
[9] W. Chen, Y. Liu, C. Yang, D. Zhu, Y. Li, (SiCp + Ti)/7075Al hybrid composites with friction stir processing, Scr. Mater. 53 (2005) 341–345.
high strength and large plasticity fabricated by squeeze casting, Mater. Sci. Eng. A [42] C.J. Hsu, P.W. Kao, N.J. Ho, Intermetallic-reinforced aluminum matrix compos-
609 (2014) 250–254. ites produced in situ by friction stir processing, Mater. Lett. 61 (2007)
[10] Y. Zhang, G. Wu, W. Liu, L. Zhang, S. Pang, Y. Wang, W. Ding, Effects of processing 1315–1318.
parameters and Ca content on microstructure and mechanical properties of squeeze [43] L. Ke, C. Huang, L. Xing, K. Huang, Al–Ni intermetallic composites produced in situ by
casting AZ91–Ca alloys, Mater. Sci. Eng. A 595 (2014) 109–117. friction stir processing, J. Alloys Compd. 503 (2010) 494–499.
[11] H. Zhang, J. Wu, Y. Zhang, J. Li, X. Wang, Y. Sun, Mechanical properties of diamond/Al [44] I.S. Lee, P.W. Kao, N.J. Ho, Microstructure and mechanical properties of Al–Fe in situ
composites with Ti-coated diamond particles produced by gas-assisted pressure in- nanocomposite produced by friction stir processing, Intermetallics 16 (2008)
filtration, Mater. Sci. Eng. A 626 (2015) 362–368. 1104–1108.
[12] D. Das, P. Mishra, S. Singh, R. Thakur, Properties of ceramic-reinforced aluminium [45] Q. Zhang, B.L. Xiao, D. Wang, Z.Y. Ma, Formation mechanism of in situ Al3Ti in Al ma-
matrix composites — a review, Int. J. Mech. Mater. Eng. 9 (2014) 1–16. trix during hot pressing and subsequent friction stir processing, Mater. Chem. Phys.
[13] F. Khodabakhshi, H. Ghasemi Yazdabadi, A.H. Kokabi, A. Simchi, Friction stir welding 130 (2011) 1109–1117.
of a P/M Al–Al2O3 nanocomposite: Microstructure and mechanical properties, [46] Q. Zhang, B.L. Xiao, Z.Y. Ma, Mechanically activated effect of friction stir processing
Mater. Sci. Eng. A 585 (2013) 222–232. in Al–Ti reaction, Mater. Chem. Phys. 139 (2013) 596–602.
[14] C. Selcuk, S. Bond, P. Woollin, Joining processes for powder metallurgy parts: a re- [47] C.J. Hsu, C.Y. Chang, P.W. Kao, N.J. Ho, C.P. Chang, Al–Al3Ti nanocomposites produced
view, Powder Metall. 53 (2010) 7–11. in situ by friction stir processing, Acta Mater. 54 (2006) 5241–5249.
[15] Z.Y. Liu, B.L. Xiao, W.G. Wang, Z.Y. Ma, Singly dispersed carbon nanotube/aluminum [48] B. Liu, Y. Liu, 27 – Powder Metallurgy Titanium Aluminide Alloys, in: M.Q.H. Froes
composites fabricated by powder metallurgy combined with friction stir processing, (Ed.), Titanium Powder Metallurgy, Butterworth-Heinemann, Boston 2015,
Carbon 50 (2012) 1843–1852. pp. 515–531.
[16] L. Lü, M.O. Lai, W. Liang, Magnesium nanocomposite via mechanochemical milling, [49] Y. Watanabe, H. Eryu, K. Matsuura, Evaluation of three-dimensional orientation of
Compos. Sci. Technol. 64 (2004) 2009–2014. Al3Ti platelet in Al-based functionally graded materials fabricated by a centrifugal
[17] R. Pérez-Bustamante, F. Pérez-Bustamante, I. Estrada-Guel, L. Licea-Jiménez, M. casting technique, Acta Mater. 49 (2001) 775–783.
Miki-Yoshida, R. Martínez-Sánchez, Effect of milling time and CNT concentration [50] S.H. Wang, P.W. Kao, The strengthening effect of Al3Ti in high temperature deforma-
on hardness of CNT/Al2024 composites produced by mechanical alloying, Mater. tion of Al–Al3Ti composites, Acta Mater. 46 (1998) 2675–2682.
Charact. 75 (2013) 13–19. [51] S.H. Wang, P.W. Kao, C.P. Chang, The strengthening effect of Al3Ti in ultrafine
[18] S. Sivasankaran, K. Sivaprasad, R. Narayanasamy, P.V. Satyanarayana, X-ray peak grained Al–Al3Ti alloys, Scr. Mater. 40 (1999) 289–295.
broadening analysis of AA 6061100-x-x wt.% Al2O3 nanocomposite prepared by me- [52] S.S. Nayak, S.K. Pabi, D.H. Kim, B.S. Murty, Microstructure-hardness relationship of
chanical alloying, Mater. Charact. 62 (2011) 661–672. Al–(L12)Al3Ti nanocomposites prepared by rapid solidification processing, Interme-
[19] R.S. Mishra, Preface to the Viewpoint Set on friction stir processing, Scr. Mater. 58 tallics 18 (2010) 487–492.
(2008) 325–326. [53] M. Kobashi, N. Inoguchi, N. Kanetake, Powder size effect on cell morphology of com-
[20] R.S. Mishra, Z.Y. Ma, Friction stir welding and processing, Mater. Sci. Eng. R 50 bustion synthesized porous Al3Ti/Al composite, Intermetallics 18 (2010)
(2005) 1–78. 1102–1105.
[21] R.S. Mishra, Z.Y. Ma, I. Charit, Friction stir processing: a novel technique for fabrica- [54] Q. Zhang, B.L. Xiao, P. Xue, Z.Y. Ma, Microstructural evolution and mechanical prop-
tion of surface composite, Mater. Sci. Eng. A 341 (2003) 307–310. erties of ultrafine grained Al3Ti/Al–5.5Cu composites produced via hot pressing and
[22] Z.Y. Ma, Friction stir processing technology: a review, Metall. Mater. Trans. A 39 subsequent friction stir processing, Mater. Chem. Phys. 134 (2012) 294–301.
(2008) 642–658. [55] F. Khodabakhshi, A.P. Gerlich, A. Simchi, A.H. Kokabi, Cryogenic friction-stir process-
[23] D.K. Lim, T. Shibayanagi, A.P. Gerlich, Synthesis of multi-walled CNT reinforced alu- ing of ultrafine-grained Al–Mg–TiO2 nanocomposites, Mater. Sci. Eng. A 620 (2015)
minium alloy composite via friction stir processing, Mater. Sci. Eng. A 507 (2009) 471–482.
194–199. [56] F. Khodabakhshi, A.P. Gerlich, A. Simchi, A.H. Kokabi, Hot deformation behavior of an
[24] A. Shafiei-Zarghani, S.F. Kashani-Bozorg, A. Zarei-Hanzaki, Microstructures and me- aluminum-matrix hybrid nanocomposite fabricated by friction stir processing,
chanical properties of Al/Al2O3 surface nano-composite layer produced by friction Mater. Sci. Eng. A 626 (2015) 458–466.
stir processing, Mater. Sci. Eng. A 500 (2009) 84–91. [57] F. Khodabakhshi, A. Simchi, A.H. Kokabi, A.P. Gerlich, M. Nosko, Effects of post-
[25] R. Bauri, D. Yadav, G. Suhas, Effect of friction stir processing (FSP) on microstructure annealing on the microstructure and mechanical properties of friction stir processed
and properties of Al–TiC in situ composite, Mater. Sci. Eng. A 528 (2011) 4732–4739. Al–Mg–TiO2 nanocomposites, Mater. Des. 63 (2014) 30–41.
[26] K.J. Hodder, H. Izadi, A.G. McDonald, A.P. Gerlich, Fabrication of aluminum–alumina [58] F. Khodabakhshi, A. Simchi, A.H. Kokabi, M. Nosko, F. Simanĉik, P. Švec, Microstruc-
metal matrix composites via cold gas dynamic spraying at low pressure followed by ture and texture development during friction stir processing of Al–Mg alloy sheets
friction stir processing, Mater. Sci. Eng. A 556 (2012) 114–121. with TiO2 nanoparticles, Mater. Sci. Eng. A 605 (2014) 108–118.
114 F. Khodabakhshi et al. / Materials Characterization 108 (2015) 102–114

[59] F. Khodabakhshi, A. Simchi, A.H. Kokabi, M. Sadeghahmadi, A.P. Gerlich, Reactive [69] A.G. Rao, K.R. Ravi, B. Ramakrishnarao, V.P. Deshmukh, A. Sharma, N. Prabhu, B.P.
friction stir processing of AA 5052–TiO2 nanocomposite: process–microstructure– Kashyap, Recrystallization phenomena during friction stir processing of hypereutec-
mechanical characteristics, Mater. Sci. Technol. 31 (2015) 426–435. tic aluminum–silicon alloy, Metall. Mater. Trans. A 44 (2013) 1519–1529.
[60] F. Khodabakhshi, A. Simchi, A. Kokabi, M. Nosko, P. Švec, Strain rate sensitivity, work [70] F.J. Humphreys, M. Haterly, Recrystallization and Related Annealing Phenomena,
hardening, and fracture behavior of an Al–Mg TiO2 nanocomposite prepared by fric- 2nd ed. Elsevier, Oxford, 2004.
tion stir processing, Metall. Mater. Trans. A 45 (2014) 4073–4088. [71] T.R. McNelley, S. Swaminathan, J.Q. Su, Recrystallization mechanisms during friction
[61] F. Khodabakhshi, A. Simchi, A.H. Kokabi, A.P. Gerlich, M. Nosko, Effects of stored stir welding/processing of aluminum alloys, Scr. Mater. 58 (2008) 349–354.
strain energy on restoration mechanisms and texture components in an [72] K. Oh-Ishi, A.P. Zhilyaev, T.R. McNelley, A microtexture investigation of recrystalliza-
aluminum–magnesium alloy prepared by friction stir processing, Mater. Sci. Eng. tion during friction stir processing of As-cast NiAl bronze, Metall. Mater. Trans. A 37
A 642 (2015) 204–214. (2006) 2239–2251.
[62] F. Khodabakhshi, A. Simchi, A.H. Kokabi, P. Švec, F. Simančík, A.P. Gerlich, Effects of [73] A.H. Ammouri, G. Kridli, G. Ayoub, R.F. Hamade, Relating grain size to the Zener–
nanometric inclusions on the microstructural characteristics and strengthening of Hollomon parameter for twin-roll-cast AZ31B alloy refined by friction stir process-
a friction-stir processed aluminum–magnesium alloy, Mater. Sci. Eng. A 642 ing, J. Mater. Process. Technol. 222 (2015) 301–306.
(2015) 215–229. [74] D. Holt, M. Jaffe, N.L. Hancox, B. Harris, H – Halpin–Tsai Equations, in: A. Kelly (Ed.),
[63] M. Gao, S. Mei, X. Li, X. Zeng, Characterization and formation mechanism of laser- Concise Encyclopedia of Composite Materials, Pergamon, Oxford 1994, pp. 125–146.
welded Mg and Al alloys using Ti interlayer, Scr. Mater. 67 (2012) 193–196. [75] M. Nakamura, K. Kimura, Elastic constants of TiAl3 and ZrAl3 single crystals, J. Mater.
[64] F.Y. Zhang, M.F. Yan, Y. You, C.S. Zhang, H.T. Chen, Prediction of elastic and electronic Sci. 26 (1991) 2208–2214.
properties of cubic Al18Ti2Mg3 phase coexisting with Al3Ti in Al–Ti–Mg system, [76] D.J. Lloyd, Particle reinforced aluminium and magnesium matrix composites, Int.
Physica B 408 (2013) 68–72. Mater. Rev. 39 (1994) 1–23.
[65] A.C. Somasekharan, L.E. Murr, Microstructures in friction-stir welded dissimilar [77] Z. Zhang, D.L. Chen, Contribution of Orowan strengthening effect in particulate-
magnesium alloys and magnesium alloys to 6061-T6 aluminum alloy, Mater. reinforced metal matrix nanocomposites, Mater. Sci. Eng. A 483–484 (2008)
Charact. 52 (2004) 49–64. 148–152.
[66] Q. Zhang, B.L. Xiao, Z.Y. Ma, In situ formation of various intermetallic particles in Al– [78] Z. Zhang, D.L. Chen, Consideration of Orowan strengthening effect in particulate-
Ti–X(Cu, Mg) systems during friction stir processing, Intermetallics 40 (2013) reinforced metal matrix nanocomposites: A model for predicting their yield
36–44. strength, Scr. Mater. 54 (2006) 1321–1326.
[67] K. Manigandan, T.S. Srivatsan, T. Quick, Influence of silicon carbide particulates on [79] A. Sanaty-Zadeh, Comparison between current models for the strength of
tensile fracture behavior of an aluminum alloy, Mater. Sci. Eng. A 534 (2012) particulate-reinforced metal matrix nanocomposites with emphasis on consider-
711–715. ation of Hall–Petch effect, Mater. Sci. Eng. A 531 (2012) 112–118.
[68] M. Karbalaei Akbari, H.R. Baharvandi, K. Shirvanimoghaddam, Tensile and fracture
behavior of nano/micro TiB2 particle reinforced casting A356 aluminum alloy com-
posites, Mater. Des. 66 (2015) 150–161.

You might also like