1 s2.0 S0925838821008227 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Alloys and Compounds 870 (2021) 159413

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Heterogeneous microstructure and deformation behavior of an


automotive grade aluminum alloy ]]
]]]]]]
]]

⁎ ⁎
S.S. Dash a, D.J. Li b, , X.Q. Zeng b, D.L. Chen a,
a
Department of Mechanical and Industrial Engineering, Ryerson University, Toronto, Ontario M5B 2K3, Canada
b
State Key Laboratory of Metal Matrix Composites, School of Materials Science and Engineering, Shanghai Jiao Tong University, 800 Dongchuan Road, Shanghai
200240, China

a r t i cl e i nfo a bstr ac t

Article history: Aluminum alloy is today considered as a prime selection for manufacturing lightweight structural com­
Received 2 January 2021 ponents in the automotive industry to increase fuel efficiency and reduce harmful emissions. The aim of this
Received in revised form 23 February 2021 study was to identify the effect of microstructure and strain rate on the tensile deformation behavior of a
Accepted 3 March 2021
high-pressure die-cast Silafont®-36 alloy, with special attention to strain hardening behavior and de­
Available online 5 March 2021
formation mechanisms. The alloy consisted of randomly oriented primary α-Al phase and Al–Si eutectic
structure in a form of heterogeneous microstructures, with Sr-modified Si particles exhibiting a coral-like
Keywords:
Cast aluminum alloy fibrous network. The local misorientations in most primary α-Al grains was below 1°, suggesting strain-free
Heterogeneous microstructure grains along with some extent of near-boundary residual strains due to the thermal mismatch between
Tensile properties aluminum and silicon. A superior strength-ductility combination was achieved, along with enhanced
Strain rate effect Young’s modulus and quality index owing to the unique heterogeneous microstructures. The cast alloy
Strain hardening behavior exhibited a smooth deformation characteristic with good coordination deformation and strong strain
hardening capacity. Strain hardening exponents evaluated via the equations proposed by Ludwik,
Hollomon, Swift, and Afrin et al., respectively, showed basically the absence of strain-rate effect from
1 × 10−5 to 1 × 10−2 s−1. During the tensile deformation, crack initiated from the sample surface and pro­
pagated through the alternate microconstituents of softer primary α-Al phase and harder eutectic structure.
© 2021 Elsevier B.V. All rights reserved.

1. Introduction poor tensile and fatigue properties. High pressure die casting
(HPDC), along with several modifications to the design of die, gating
Due to their lightweighting trait along with other advantages like system, venting and vacuum systems, and die-lubrication, can ef­
high specific strength, superior corrosion resistance, better ductility, fectively eliminate the critical gas entrapment in the casting process
and high thermal and electrical conductivity, the application of of Al alloys [10–14]. Such die-cast Al alloys contain Si as a primary
aluminum alloys in North American automobiles is expected to rise alloying element, aiming to increase fluidity, ensure proper die
12% by 2026, bringing the average aluminum content per vehicle filling, and prevent mold wall sticking along with high machinability
from 208 kg (2020) to 233 kg (2026) [1]. While both cast and and high-integrity. These Al–Si alloys are sometimes referred to as
wrought aluminum alloys are used to manufacture a variety of au­ “silumins”, with hypoeutectic and near-eutectic Al–Si alloys
tomotive components, on the industrial scale cast Al alloys are more (9–12.67 wt% Si) being preferred for producing major vehicle com­
popular due to lower production cost. Manufacturing processes such ponents like engine blocks, wheels, main chassis, door panels and
as pressure die casting, permanent mold casting, thixocasting, frames, etc. Silafont®-36 Al–Si alloy containing very low Fe content,
rheocasting, etc., are able to produce near-net-shape components at developed by Rheinfelden Alloys GmbH, was one of the first HPDC
a larger scale, which help reduce the production time and increase alloys which had higher elongation and superior tensile and fatigue
efficiency [2–9]. However, the presence of some imperfections strength [11].
during casting like the entrapment of oxides, gas and shrinkage Mechanical properties of cast Al–Si alloys are dependent on the
pores, facilitate the initiation of cracks in these alloys, leading to as-cast microstructures involving the morphology of Si particles
(size, shape, and distribution), primary dendrite parameters, and
grain size [5]. These microstructural features change with processing

Corresponding authors. parameters, solidification rate, chemical modification, heat
E-mail addresses: lidejiang@sjtu.edu.cn (D.J. Li), dchen@ryerson.ca (D.L. Chen).

https://doi.org/10.1016/j.jallcom.2021.159413
0925-8388/© 2021 Elsevier B.V. All rights reserved.
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Table 1 microstructure, focusing on the strain hardening behavior and me­


Chemical composition (wt%) of the present as-cast Silafont®-36 alloy. chanisms along with the effect of strain rate.
Si Fe Cu Mn Mg Zn Ti Sr Al

10.34 0.091 0.0008 0.62 0.32 <0.0010 0.066 0.013 Balance 2. Material and experimental procedure

In the present study, Silafont®-36 alloy ingots were purchased


from a company in China authorized by Rheinfelden Alloys GmbH.
treatment, and thermomechanical processes [15]. Silafont®-36 alloy Molten Silafont®-36 alloy maintained at a temperature of 680 °C was
contains ~10% of Si with other minor alloying elements like Mn and poured into a preheated mold at 180 °C at a pressure of 80 MPa. The
Mg, which enhance the ductility and strength of the alloy. Si parti­ fast and slow ram speeds were kept constant at 2 m/s and 0.25 m/s,
cles, being present in the eutectic structure, could be acicular, respectively. A vacuum level of about 100–150 mbar was applied in
elongated or spherical morphologies. These shapes affect the stress the die-casting process. The chemical composition of the alloy after
field interaction during deformation, with a higher stress con­ casting was determined via plasma mass spectrometry and is pre­
centration being present for acicular and elongated shapes in com­ sented in Table 1. Being a low Fe-containing die-cast aluminum
parison with spherical shapes [16]. The minute addition of Sr to even alloy, it was designed to improve the elongation and strength with
ppm levels have been reported to lead to the modification of Si subsequent Mn addition, comparable to other die-cast aluminum
morphology to more rounded-shapes, which increased the ductility alloys. The alloy contains 10.34% Si as a major alloying element,
[17–21]. The addition of Mg in Al–Si alloys leads to the formation of making it a hypoeutectic Al–Si alloy. The dogbone-shaped rectan­
Mg2Si phase, which is known for its precipitation strengthening gular and cylindrical samples following ASTM E8 standards were
effect in cast Al–Si–Mg alloys [22]. Zovi and Casarotto [10] studied directly cast using a specifically designed mould, as shown in
the effect of alloying elements on the mechanical properties of as- Fig. 1(a) [3]. Fig. 1(b) shows the geometry and dimensions for the
cast and heat-treated Silafont®-36 alloy. Addition of Mn helped im­ standard specimens. The rectangular (Fig. 1(i)) dog-bone samples
prove the strength of the as-cast alloy, and reduce the sticking were used for the microstructural characterization to facilitate the
tendencies. However, with increasing Mn content the elongation of sample orientations indicated, and the cylindrical samples (Fig. 1(ii))
the alloy decreased due to the formation of Al–Mn–Fe–Si brittle for the tensile tests. Since the rectangular and cylindrical specimens
intermetallic phases [23]. The presence of needle-like Fe-containing were positioned symmetrically in the mould (Fig. 1(a)), their mi­
intermetallic compounds downgraded the tensile elongation and crostructure and mechanical properties would be anticipated to be
strength by acting as sources of stress concentration and crack in­ equivalent.
itiation sites. Small pieces cut from the as-cast samples using a slow-speed
Some limited studies on the mechanical behavior of Silafont®-36 diamond cutter were manually ground with abrasive SiC papers (Grit
cast alloy have been conducted. Lee and Mishra [24,25] studied the #320, #400, #600, #1200, #2000, and #4000), and then polished
quantitative features of eutectic Si particles with varying cooling using diamond paste and colloidal silica. The polished surfaces were
rates in Silafont®-36 alloy, and observed that finer grains blocked etched using Keller’s reagent (5 ml HNO3, 3 ml HCl, and 2 ml HF in
more dislocations produced around indentation deformation, re­ 190 ml distilled water) for ~3 min. The same samples were then
sulting in higher hardness values. Faster cooling rates also increased deeply etched for 7 min to better observe the morphology of eutectic
the solubility of alloying elements in the primary α-Al phase due to a phases. Microstructures were observed via JSM-6380LV scanning
higher extent of supersaturation, leading to stronger solid solution electron microscope (SEM) attached with Oxford energy-dispersive
strengthening. Refined eutectic Si particles in the case of Sr or spectroscopy (EDS) detector. EDS was carried out on the sample
phenyl-TSP (trisilanol phenyl polyhedral silsesquioxane) modified surfaces in both line scan and elemental mapping for the as-cast
Al–10%Si–Mn–Mg cast alloy, resulted in a significant increase in microstructures. Electron backscatter diffraction (EBSD) examina­
elongation [16]. Yuan et al. [26] studied the effect of solution time on tions to identify the as-cast grain orientations and phases were also
the tensile properties of a similar Al–10%Si–Mn–Mg HPDC alloy. The conducted at a step-size of 0.1 µm. Image analysis software (ImageJ)
available studies mainly involved the effect of casting parameters. was used to determine the volume fraction of each phase. X-ray
There has been no information about the quality index of the Sila­ diffraction was performed using a PANalytical X-ray diffractometer
font®-36 cast alloy (or how high the value should be for achieving (XRD) to identify the phases present in the as-cast Al–Si alloy. Cu Kα
effective casting). It is also unclear how the strain hardening behaves X-ray radiation with a wavelength λ = 0.15406 nm with a setting of
and if strain rate sensitivity is present in the Silafont®-36 alloy 45 kV and 40 mA was used. The diffraction angle (2θ) range was set
especially with a heterogeneous microstructure. The objective of this to vary from 20°–90° at a scanning rate of 0.05° s−1.
study is, therefore, to investigate the tensile deformation behavior of Tensile tests were performed according to ASTM E8M standards
Silafont®-36 alloy in relation to the as-cast heterogeneous at room temperature on a computerized United testing system. The

Fig. 1. (a) High-pressure die-casting mould used for the Silafont®-36 alloy [3], showing (i) rectangular and (ii) cylindrical samples positioned symmetrically, and (b) schematic
diagram for the specimen geometry and dimensions (in mm) for (i) and (ii).

2
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

tests were conducted on cylindrical samples with a gauge length of beneficial for acquiring superior mechanical properties in the cast
50 mm at four different strain rates of 1 × 10−5, 1 × 10−4, 1 × 10−3, and alloys [26,27]. This kind of structural modification is due to the
1 × 10−2 s−1. At least two samples were tested at each strain rate, with presence of minute amount of strontium (Sr) in the alloy
only mild grinding to remove the burrs after demolding. Yield [11,17,28–30]. The morphology of Si, which can be modified due to
strength (YS), ultimate tensile strength (UTS), elongation (%EL), the presence of alloying elements like Sr or other modifiers like Sc
quality index (Q), and strain hardening exponent (n) were evaluated or Na, plays an important part in shaping the mechanical properties
from the obtained stress-strain curves. Some typical fracture of die-cast alloys. As a result of such elongated Si particles, the
surfaces were examined via SEM to identify the tensile fracture elongation of the alloys gets significantly reduced [26]. Some mi­
mechanisms in this low-iron die-cast Al–Si alloy. crostructures after deep-etching show the presence of externally
solidified crystals (ESCs), which denote the primary α-Al solidifi­
3. Results and discussion cation taking place in shot sleeve. As-cast microstructures could
also contain some minute content of β-Mg2Si phase sitting on the
3.1. Microstructure Si eutectic particles [22,26,31,32], due to the stronger chemical
affinity between the atoms of Mg and Si [33].
Fig. 2 shows typical microstructures of the as-cast Silafont®-36 Primary α-Al dendrites in the Silafont®-36 alloy that formed
alloy, which were viewed from the longitudinal direction (LD). Si­ through dendritic solidification during casting, do not have long
milar microstructural features were also observed from the normal arms and are quite fragmented (Fig. 2). From (Fig. 2(d)), it can be
direction (ND) and transverse direction (TD), which indicated no seen that the finer and fibrous structure of eutectic Si particles is due
obvious anisotropy along different directions. It should be noted to modification by the trace addition of 130 ppm of Sr. The Sr gets
that both the rectangular and cylindrical specimens have a typical absorbed onto the surface of Si, as verified by the EDS line scan
die casting microstructure from the surface to center, i.e., a gradient analysis shown in Fig. 3(a), where both Si and Sr demonstrated a
microstructure (ultra-fine grains in the skin and coarse grains in higher elemental distribution in the areas of eutectic structure while
the center). The features of primary α-Al and interlamellar Al–Si other elements like Mg, Mn and Fe were absent. Similar findings
eutectic structure become more visible with increasing magnifi­ were also reported by Timpel et al. [20] who showed that the co-
cation, as seen from Fig. 2(a)-(c). The rod-shaped particles with segregation of Sr–Al–Si together can change the stacking fault en­
light-grey color represent eutectic Si, and the dark-grey patches in- ergy. The presence of Sr can effectively slow down the growth rate of
between Si particles stand for eutectic Al, as indicated in Fig. 2(c). It eutectic Si particles relative to Al matrix. The mechanism involves
should be noted that a longer etching time led to the removal of inducing stacking faults and twinning [18,34], due to which the
some inclusions, whose imprints could be seen as the dark pore- unidirectional growth of Si particles is prevented [19,27]. The poi­
like spots on the micrographs. Fig. 2(d) shows a deep-etched mi­ soning of the Si growth sites at the advancing interface by Sr sug­
crostructure of the alloy, where the localized corrosion of eutectic gests that the eutectic modification is controlled by a growth
Al phase in the eutectic structure revealed a coral-like structure of mechanism – commonly known as the “twin plane re-entrant edge”
eutectic Si particles, which was reported by some researchers to be (TPRE) mechanism. The occurrence of poisoning retards the growth

Fig. 2. SEM micrographs of the Silafont®-36 alloy with (a–c) normal etching, and (d) deep etching, observed in the longitudinal (sample length) direction (LD), at different
magnifications in the secondary electron imaging (SEI) mode.

3
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Fig. 3. EDS line scan results obtained from (a) an etched sample surface across primary α-Al and eutectic structure, and (b) an unetched sample surface across the second phase
particles and base metal.

of Si and increases undercooling, thus forming finer particles (black lines in Fig. 4(a–d)). The absence of some obvious grain
[17,20,21]. Vandersluis et al. [27] also observed a similar effect of Sr boundaries was due to the overlapping with the eutectic structure.
on the Si particle morphology in as-cast A319 alloy with an average Fig. 4(c) shows the phase distribution containing two major phases,
aspect ratio of Si particles of 1.5–2 with Sr addition. Al and Si, and some minor presence of Mg2Si phase. Si is basically
EDS analysis was also performed on an unetched specimen, as present in the inter-dendritic regions, confirming its existence in the
shown in Fig. 3(b), which revealed the presence of Mn and Fe across eutectic structure. Fig. 4(d) shows the local misorientation or Kernel
the white hexagon-like particles, i.e., inclusions in the as-cast Si­ average misorientation (KAM) maps obtained for the Al phase in the
lafont®-36 alloy. Mohammed et al. [3] also reported the presence of as-cast Silafont®-36 alloy. KAM maps depict the extent of strain
Al8(Fe,Mn)2Si particles in a newly developed Al–5.5Mg–2.5­ distribution inside a grain [37,38]. Local misorientations of <5° were
Si–0.6Mn–0.2Fe cast aluminum alloy. However, the amount of such considered in this map, to avoid the effect of large misorientations
Fe- and Mn-containing inclusions in the present alloy is scarce due like the grain boundaries. Fig. 4(e) shows the frequency distribution
to a very low content of Fe (~0.09%, Table 1), since the presence of of the local misorientations, indicating that the maximum frequency
excess Fe-rich intermetallic compound is known to result in a is below 1°. This suggests that the as-cast microstructure consists
higher stress concentration and cause early cracking and low mostly of strain-free regions, but with some extent of internal strain
ductility. Addition of Mn (0.62%) in the alloy helps prevent sticking mainly near the boundaries. This was likely due to the pretty large
between the mold wall and casting, but the presence of excess Mn- thermal mismatch between the Al and Si during fast cooling after
rich inclusions can also lead to brittle fracture and decrease the casting (i.e., from the eutectic temperature of 577 °C to room tem­
ductility, which will be discussed in the later section of tensile perature) like the situation of particulate-reinforced metal matrix
properties. composites, where the coefficient of thermal expansion (CTE) is
~23.6 × 10−6 °C−1 for Al, and ~2.6 × 10−6 °C−1 for Si with a large dif­
3.2. EBSD analysis ference ΔCTE of ~21 × 10−6 °C−1. The presence of areas with high KAM
values inside some primary α-Al grains is most likely due to a certain
EBSD results of the as-cast microstructure observed from the level of dislocations generated in the casting process, which follows
normal direction (ND), provided some interesting features as shown several mechanisms like merging dendritic grains, stress build-up at
in Fig. 4(a–d). The color changes in Fig. 4(a) revealed random tex­ particle-matrix interface, and thermal and shrinkage stresses
tures in the primary α-Al dendrites. The primary Al phase in an A413 [37,39]. This also corroborates the fact that the KAM values are
cast aluminum alloy was also observed to exhibit varying orienta­ usually high at the Si particle and Al matrix interface.
tions [35]. The grain boundary-like network of white-colored regions
in Fig. 4(a) included eutectic Si particles along with a small amount 3.3. X-ray diffraction
of Mg2Si phase, as shown in Fig. 4(b). It should be noted that pseudo-
symmetric mis-indexing would be present due to the insufficient X-ray diffraction (XRD) pattern of the as-cast Silafont®-36 alloy is
resolution and the low quality of indexing in the fine lamellar eu­ shown in Fig. 5. Distinct peaks of α-Al and Si in the eutectic structure
tectic region. Similar EBSD analyses were conducted by Chankit­ are observed, along with the presence of Mg2Si and AlFeMnSi
munkong et al. [36] for an Al–12%Si–Cu alloy, where finer eutectic (Al9Fe0.84Mn2.16Si) phases as seen from the minor peaks. The pre­
structures were also difficult to index. Boundaries with misorienta­ sence of a small amount of Mg2Si corresponds well to the EBSD
tion angles between 2°–15° were considered as low angle grain analysis (Fig. 4(b and c)). Indeed, Mg2Si is one of the most commonly
boundaries (LAGBs), whereas boundaries with misorientation angles observed intermetallic phases in many Al–Mg–Si alloys [3,22,39,40].
of >15° were considered as high angle grain boundaries (HAGBs) Since Mg2Si with a cubic lattice (Fm3m, 12 atoms/unit cell, lattice

4
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Fig. 4. EBSD orientation maps of (a) α-Al phase and (b) Si and Mg2Si in the eutectic structure, (c) EBSD phase distribution map of α-Al, Si and Mg2Si phases, (d) local misorientation
or KAM (Kernel Average Misorientation) map for Al-phase, with the corresponding (e) local misorientation distribution in the as-cast Silafont®-36 alloy.

parameter a = 0.635–0.640 nm) has a melting temperature of from the fracture surface. Similar intermetallic particles have also
1087 °C, a density of 1.88 g/cm3, and a Vickers hardness of 4.5 GPa, it been reported in several Al–Si–Mn alloys containing trace amounts
can act as an effective lightweight strengthening phase, leading to of Fe [42–44].
higher hardness, tensile strength and fatigue life of Al alloys
[3,39,41]. In an as-cast Al–Mg–Si alloy with a Mg-to-Si ratio of 3.4. Tensile properties and enhanced Young’s modulus
roughly 2:1, Mg2Si phase was observed to be extensively present
with the microstructure containing nearly ideal eutectic layers of Fig. 6(a) shows a typical engineering stress-strain curve of as-cast
Mg2Si [3]. As seen from Fig. 5, some weak peaks corresponding to an Silafont®-36 alloy tested at a strain rate of 1 × 10−4 s−1. It exhibits a
AlFeMnSi (Al9Fe0.84Mn2.16Si) intermetallic compound could also be typical smooth flow curve with gradual transition from elastic to
observed. This intermetallic compound corresponds well to the Mn- plastic deformation without the occurrence of yield-point phe­
rich particles observed from the EDS analysis (Fig. 3(b)) and later nomenon. The Young’s modulus in the initial elastic deformation

5
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Fig. 5. X-ray diffraction pattern showing the phases and intermetallic compounds in the as-cast Silafont®-36 alloy.

Fig. 6. (a) A typical engineering stress-strain curve tested at a strain rate of 1 × 10−4 s−1, and (b) the yield strength (YS), ultimate tensile strength (UTS), elongation (%EL), and
quality index (Q), along with (c) hardening capacity (Hc) and (d) strain hardening exponent (n) (evaluated using the equations of Ludwik [70], Hollomon [71], Swift [72], and Afrin
et al. [67]) for the as-cast Silafont®-36 alloy as a function of strain rate.

6
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Table 2
Comparison of the tensile properties of the present as-cast Silafont®-36 (Al–10.34Si–0.62Mn–0.32Mg) alloy with those of other similar or common as-cast Al alloys.

Alloy Yield Ultimate tensile Elongation, % Strain hardening Hardening capacity (Hc)
stress, MPa strength, MPa exponent (n)

Al–10.34Si–0.62Mn–0.32Mg (present study) 169 ± 6 314 ± 3 6.2 ± 1.8 0.26 ± 0.01 0.86 ± 0.06
A319 [56] 130 235 2.5 – 0.81
A380 [57] 160 325 3.5 – 1.03
Al–Mg–Si–Mn [57] 183 324 8.3 – 0.77
Al–5.5Mg–2.5Si–0.6Mn–0.2Fe [3] 185 304 6.3 0.25 0.72
A356 [58] 90 185 8 – –
Al–11.6Si [59] – 180 2 – –
Al–5Si [60] – 209 ± 10.1 4.5 ± 1.2 – –
Al–8Si [60] – 276 ± 4.5 6.0 ± 0.4 – –
Al–18Si [60] – 249 ± 8.3 2.2 ± 0.2 – –
Al–8.94Si–0.93Fe–3.24Cu [61] 114 135 1.52 – –

(EA) could be estimated via the following upper (u) and lower (l) without heat treatment is mainly due to a lower amount of Fe
bound expressions of rule of mixtures, (Table 1, as mentioned earlier) along with an optimized composition
of Mn. Wu et al. [7] observed the deterioration of tensile properties
EA (u) = EAl × VAl + ESi × VSi, (1)
with increase in Mg content beyond 0.45%, that helped in stabilising
EAl × ESi the Fe-based intermetallic in Al–7%Si–Mg alloy microstructure. Zovi
EA (l) = , and Casarotto [10] studied the effect of Mn amount on the mechanical
VAl × ESi + VSi × EAl (2)
properties of Silafont®-36 alloy, and found that at about 0.5–0.8% Mn
where VAl and VSi are the volume fraction of pure Al and Si (VAl + VSi = 1), content produces an elongation more than 8%. Table 2 summarizes
and EAl and ESi are the Young’s modulus of Al and Si, respectively. The the tensile properties of several common automotive Al alloys in the
average value of VSi is ~0.22 via image analysis. Using the EAl as 69 GPa as-cast condition [3,56–61]. It is seen that the present as-cast Sila­
and ESi as 162 GPa [45], the Young’s modulus of the alloy obtained from font®-36 alloy lies at the high end of as-cast Al alloys with a superior
Eqs. (1) and (2) lies in-between ~78 and ~88 GPa. The average value combination of YS, UTS and %EL.
estimated from the linear region of tensile stress-strain curves was A parameter known as quality index (Q, in MPa) by considering
~77 GPa, which is close to the lower bound calculated via the rule of both UTS and elongation has been used to reflect the quality of as-
mixtures. This suggests superior bonding between Si particles and Al cast alloys. It is defined as follows [62],
matrix in the present as-cast Silafont®-36 alloy due to the application
Q = UTS (inMPa) + log(Elongation (in %)), (3)
of a high pressure (80 MPa) during casting. Several recent studies have
shown that the increase of Si as an alloying element from 7% to 12% in where δ was taken to be a constant (=150 MPa) for the original
cast Al alloys leads to an increase of Young’s modulus from 73 to Al–7Si–0.4Mg alloy and assumed to be the same for the current alloy
77 GPa, as the higher Si content results in the formation of more as well. The parameter was proposed to show a compromise of the
covalent bonds. This phenomenon is present in many hypo- and hyper- UTS and ductility of the cast alloys and assess the optimal conditions
eutectic alloys [46–48]. Mandal and Makhlouf [49] also showed the for castings. The calculated Q values versus strain rates are shown in
beneficial effect of HPDC process, along with the presence of Mg2Si Fig. 6(b), lying in-between 400 and 470 MPa and reflecting the
phase, in increasing the Young’s modulus of a hypereutectic Al–Si alloy. uniformity of the present as-cast alloy with respect to the quality of
Fig. 6(b) shows the values of YS, UTS, quality index (Q), and %EL, casting. The obtained Q values are higher than that (~300 MPa) of an
respectively, as a function of strain rate. It is seen that a YS of as-cast Al–10.8%Si alloy reported by Mohamed et al. [63]. Several
160–180 MPa and UTS of 310–320 MPa are achieved in the present other cast Al alloys developed without using HPDC process [64–66]
as-cast Silafont®-36 alloy, which are higher than the specified exhibited relatively lower Q values than the present alloy, which
YS = 120–150 MPa and UTS = 250–290 MPa, while the obtained %EL demonstrates the efficiency of this process in manufacturing high-
of 5–10% also meets the specified range of 5–11%. The strain rate end automotive components.
sensitivity is seen to be basically absent, since the obtained YS, UTS
and %EL values are roughly the same at the applied strain rates 3.5. Strain hardening behavior
ranging from 1 × 10−5 s−1 to 1 × 10−2 s−1 within the experimental
scatter. Similarly, Zhang et al. [50] studied the strain rate de­ As seen from Fig. 6(a), the strain hardening of the present as-cast
pendency of tensile properties of an Al–5%Mg–0.6%Mn alloy, and Silafont®-36 alloy after yielding is pretty smooth and uniform, which
also observed almost no changes in the YS and UTS. The enhanced could be better characterized using hardening capacity, hardening
mechanical properties of the as-cast Silafont®-36 alloy would be exponent, and hardening rate. The hardening capacity (Hc) can be
directly related to the composite-like heterogeneous micro­ defined as the following normalized ratio between the UTS (σUTS)
structures as seen from Fig. 2 and also reported in [51,52], along with and YS (σy) [67],
the presence of some well-dispersed β-Mg2Si and Al9Fe0.84Mn2.16Si UTS y UTS
particles (Figs. 3(b), 4 and 5). The unique heterogeneity of this alloy HC = = 1.
y y (4)
is hence attributed to the presence of intermetallic particles and Si in
the dendritic structure and can also be compared to several auto­ The evaluated hardening capacity as a function of strain rate for
motive Al alloys and particulate-reinforced Al-based composites the as-cast Silafont®-36 alloy is shown in Fig. 6(c). It is seen that like
[53–55]. A combined role of the optimal process control with a YS, UTS, and %EL in Fig. 6(b), the hardening capacity of this alloy is
proper addition of alloying elements, mainly 10.34% Si, 0.32% Mg, also nearly independent of the applied strain rate. The hardening
0.62% Mn and 0.091% Fe (Table 1), leads to such a synergic effect. capacity of this alloy is observed to be higher than several other Al
Industrial components with high operating stresses against impact alloys (Table 2), except for A380 alloy [57]. This is likely due to the
have a Mg content at levels of 0.24–0.35% which help improve the lower YS of A380 alloy which leads to a higher dislocation storage
strength [10,11]. The larger elongation obtained in the cast alloy capacity during plastic deformation and hence a higher hardening

7
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

capacity, compared to Silafont®-36 alloy [67]. Similar results on a Extending this mechanism to a heterogeneous material, in order to
weak strain-rate dependent hardening capacity were also reported accommodate the strain difference between the softer phase (Al)
for magnesium alloys in [68,69]. and the harder phase (eutectic Si), GNDs are generated making the
Strain hardening behavior could further be characterized via softer phase stronger and globally increasing the yield strength of
strain hardening exponent. Several equations related to the strain the material. Generation of back stress due to the presence of GNDs
hardening exponent have been proposed. The first equation pro­ results in an enhancement of the strength-ductility of these alloys.
posed by Ludwik [70] can be expressed as follows, Eq. (9) could be further modified as follows by including the terms of
n1, dislocation density caused by statistically stored dislocations (ρS)
= y + K1 (5)
and GNDs (ρG) [51],
where n1 is the strain hardening exponent, K1 is the strength coef­
ficient that represents a strength increment corresponding to ε = 1, σy y S + G . (10)
is the yield strength, and σ and ε are the true stress and true strain,
The hardening effect due to the involvement of GND back-stress
respectively. A simplified equation proposed by Hollomon [71] is,
has been coined as hetero-deformation induced (HDI) strengthening
= K n, (6) in the heterogeneous microstructures. Some recent studies on the
where n is the strain hardening exponent and K is the strength eutectic multi-principal elemental alloys (EMPEAs), nanocrystalline
coefficient. Later, Swift [72] proposed the following modified and ultra-fine grained structured alloys have also shown the benefits
equation, of designing the alloys based on heterogeneous micro­
structures [76–79].
= KS ( + n
y ) S, (7)
where nS and KS are strain hardening exponent and strength coef­ 3.6. Fractography
ficient, respectively, and εy is the yield strain. Afrin et al. [67] further
modified the strain hardening exponent related equation by con­ Typical fracture surfaces of the specimens tested at high and
sidering net flow stress and net plastic strain of a material after low strain rates are shown in Fig. 7(a)–(d). The fractographs at
yielding as follows, lower magnifications in Fig. 7(a) and (c) show crack initiation from
n* , the specimen surface and typical crack flow/propagation patterns
y = K* ( y) (8)
that led to the fracture of the specimens. It is seen that strain rate
where n* is the strain hardening exponent and K* is the strength did not have a significant effect on the fracture surface char­
coefficient which reflects a strength increment corresponding to acteristics, corroborating the results shown in Fig. 6. Fig. 7(b) and
(ε – εy) = 1. Fig. 6(d) presents the strain hardening exponents (d) shows a magnified view of crack propagation region boxed in
evaluated via the above four equations, using the identical set of Fig. 7(a) and (c), respectively, where the Al–Si eutectic structure
stress-strain data for each sample in the uniform plastic deformation and primary α-Al phase are visible. The eutectic structure exhibits
region in-between YS and UTS. It is seen that Ludwik’s equation finer brittle-fracture facets arising from the broken Si particles,
gives the highest strain hardening exponent, while Hollomon’s whereas the primary α-Al phase shows more ductile character­
equation gives the lowest. The equations by Afrin et al. and Swift lie istics with tear ridges. Purcek et al. [80] described the underlying
in-between the two strain hardening exponent values. Similar re­ mechanism of crack development in an as-cast Al–Si alloy
sults on the relative positioning of these strain hardening exponent showing brittle fracture. Fig. 8(a) and (b) shows a cross-sectional
values evaluated on the basis of Eqs. (5)–(8) were also reported in view of the fractured surface at different magnifications in order
[68,69]. These strain hardening exponent values do not vary much to understand the underlying crack propagation mechanism. Since
with strain rate, hence it could also be concluded that the strain rate Si particles are more brittle than the α-Al phase, the cracking oc­
sensitivity of the Silafont®-36 alloy is basically absent based on this curs easier at low strains. As the deformation progresses, stress
point, as in the case of AZ91D cast magnesium alloys [69]. As seen concentration occurs at the tip of the Si particles or between Si
from Fig. 2, the Silafont®-36 alloy consisted of heterogeneous mi­ particles and α-Al phase, resulting in the formation of a crack that
crostructure with primary α-Al as the matrix phase, similar to the first propagates through the Si-rich eutectic structure area. Then
microstructure in a ferrite-martensite dual-phase steel [73]. Mate­ the crack grows into the primary α-Al phase with tear ridges
rials with heterogeneous microstructures, also known as hetero- (Fig. 8(a)) until reaching its boundary adjacent to next eutectic
structure materials, exhibit a strength-ductility synergy due to the structure area, then gets into this eutectic structure to continue
coordination deformation of hard and soft phases in the micro­ propagation in the heterogeneous microstructures of the Sila­
structures in different patterns [51,74], which involves the genera­ font®-36 aluminum alloy (Fig. 2). Most of the eutectic Si particles
tion of geometrically necessary dislocations (GNDs) in the soft phase get fractured in the process and result in a river-bed like feature on
in the vicinity of phase boundaries (Fig. 4(d)). Deformation in these the fractographs. Such a quasi-periodic propagation mechanism in
materials occurs in three stages, where the soft and hard phases the alternate microconstituents of relatively harder eutectic
interact differently at different strains [51]. Strain hardening of structure regions and relatively softer primary α-Al regions, as
materials in the uniform plastic deformation region depends on the schematically shown Fig. 8(c), leads to the final failure of this alloy
density of dislocation. Dislocation strain field interactions result in under increasing tensile loading, with the fractured Si particles
an increasing net flow stress of metals and are related in the form of creating some cleavage-like facets (Fig. 8(b)). A similar mechanism
the following expression [68,75], was also observed by Jiang et al. [14], which led to superior duc­
tility in the Al–Si cast alloy. The fracture surface did not contain
y , (9)
any debris of Si or intermetallic particles, suggesting that the
where ρ is the dislocation density inside the material. During plastic formation of cracks due to voids around Si particles did not occur
deformation, there is an increase in the number of dislocations in this alloy, as noted in [81,82]. Also, the AlFeMnSi particles could
stemming from the dislocation multiplication along with the for­ be identified by EDS on the fracture surface, as characterized by
mation of new dislocations. This results in a decrease in the spacing cleavage facets with hexagon-like shapes (Fig. 7(d)), corre­
between the dislocations, thereby leading to repulsive interactions. sponding nicely to the Al9Fe0.84Mn2.16Si particles detected by the
As a result, the motion of other dislocations becomes impeded and a earlier EDS and XRD analyses (Figs. 3 and 5). The superior strength
higher stress following Eq. (9) is required to continue deformation. (YS = 169 MPa and UTS = 314 MPa, Table 2) along with 5–10%

8
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Fig. 7. Typical fracture surfaces of the as-cast Silafont®-36 alloy after the tensile tests at (a and b) a higher strain rate of 1 × 10−2 s−1, and (c and d) a lower strain rate of 1 × 10−5 s−1.

ductility obtained in the as-cast Silafont®-36 alloy can be attrib­ using the equations proposed by Ludwik, Hollomon, Swift and
uted to the presence of such a heterogeneous composite-like Afrin et al., respectively, showed basically no effect of strain rate
structure (Fig. 2) along with the formation of tear ridges in the in the applied range from 1 × 10−5 to 1 × 10−2 s−1.
primary Al phase (Figs. 7 and 8). 5) Crack initiation was observed to occur from the sample surface
and crack propagated through the alternate microconstituents
4. Conclusions of relatively harder eutectic structure and softer primary α-Al
phase.
1) The primary equiaxed α-Al grains and Al–Si eutectic structure
were present in a form of heterogeneous microstructures in the
HPDC Silafont®-36 cast aluminum alloy, with Sr-modified Si CRediT authorship contribution statement
particles exhibiting a coral-like network.
2) EBSD analysis showed the as-cast microstructure with random S.S. Dash: Methodology, Investigation, Data curation, Formal ana­
orientations and the presence of Mg2Si phase along with alu­ lysis, Validation, Writing - original draft. D.J. Li: Conceptualization,
minum and silicon. The local misorientations in the primary α-Al Methodology, Resources, Writing - review & editing. X.Q. Zeng:
grains was below 1°, suggesting strain-free grains along with Conceptualization, Methodology, Resources, Funding acquisition. D.L.
some extent of near-boundary residual strains arising from the Chen: Conceptualization, Methodology, Resources, Supervision, Writing
thermal mismatch between aluminum and silicon during cooling - review & editing.
from the eutectic temperature to room temperature like a par­
ticulate-reinforced metal matrix composite. Data availability
3) A superior combination of mechanical properties was attained,
including enhanced Young’s modulus and quality index due to The raw/processed data required to reproduce these findings
the unique heterogeneous microstructures and good bonding cannot be shared at this time as the data also forms part of an on­
between Si particles and aluminum matrix owing to high pres­ going study.
sure during casting. The addition of a proper amount of Mn and
Mg alloying elements along with a low Fe content helped achieve
a YS of 169 MPa, a UTS of 314 MPa and a %EL of 5–10% in the as- Declaration of Competing Interest
cast condition.
4) This cast aluminum alloy exhibited a smooth deformation char­ The authors declare that they have no known competing fi­
acteristic with superior coordination deformation and high strain nancial interests or personal relationships that could have appeared
hardening capacity. Strain hardening exponents determined to influence the work reported in this paper.

9
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

Fig. 8. (a and b) Cross-sectional fracture surfaces showing the crack and its propagation at different magnifications, and (c) schematic representation of crack propagation through
the heterogeneous primary and eutectic structures during tensile deformation of Silafont®-36 alloy.

Acknowledgements [5] A.R. Emami, S. Begum, D.L. Chen, T. Skszek, X.P. Niu, Y. Zhang, F. Gabbianelli,
Cyclic deformation behavior of a cast aluminum alloy, Mater. Sci. Eng. A
516 (2009) 31–41, https://doi.org/10.1016/j.msea.2009.04.037
The authors would like to thank the Natural Sciences and [6] Y.H. Zhang, C.Y. Ye, Y.P. Shen, W. Chang, D.H. StJohn, G. Wang, Q.J. Zhai, Grain
Engineering Research Council of Canada (NSERC) and the National refinement of hypoeutectic Al-7wt%Si alloy induced by an Al-V-B master alloy,
Natural Science Foundation of China (No. 51825101) in the form of J. Alloy. Compd. 812 (2020) 152022, https://doi.org/10.1016/j.jallcom.2019.
152022
international research collaboration. D.J.L. and X.Q.Z. would like to [7] X. Wu, H. Zhang, Z. Ma, T. Tao, J. Gui, W. Song, B. Yang, H. Zhang,
thank the National Key Research and Development Program (No. Interactions between Fe-rich intermetallics and Mg-Si phase in Al-7Si-xMg al­
2016YFB0101704) supported by the Ministry of Science and loys, J. Alloy. Compd. 786 (2019) 205–214, https://doi.org/10.1016/j.jallcom.2019.
01.352
Technology of China and the Shanghai Science and Technology [8] W. Jiang, Z. Fan, X. Chen, B. Wang, H. Wu, Combined effects of mechanical vi­
Committee (No. 18511109302). The authors would also like to thank bration and wall thickness on microstructure and mechanical properties of A356
Messrs. S.M.A.K. Mohammed, A. Machin, and Q. Li for their assis­ aluminum alloy produced by expendable pattern shell casting, Mater. Sci. Eng. A
619 (2014) 228–237, https://doi.org/10.1016/j.msea.2014.09.102
tance and easy access to the laboratory facilities of Ryerson
[9] J. Zhu, W. Jiang, G. Li, F. Guan, Y. Yu, Z. Fan, Microstructure and mechanical
University and their assistance in the experiments. properties of SiCnp/Al6082 aluminum matrix composites prepared by squeeze
casting combined with stir casting, J. Mater. Process. Technol. 283 (2020) 116699,
References https://doi.org/10.1016/j.jmatprotec.2020.116699
[10] A. Zovi, F. Casarotto, Silafont-36, the low iron ductile die casting alloy develop­
ment and applications, Metall. Ital. 99 (2007) 33–38.
[1] Abraham A., Schultz R., Rakoto B., Murphy J., Ling L., Merta M., et al. 2020 North [11] Rheinhfelden, Primary Aluminium, Alloys for High Pressure Die Casting, CAR DS-
America Light Vehicle Aluminum Content and Outlook: Final Report Summary. Mould Des, 2015, pp. 1–68.
2020. [12] J.S. Shin, S.H. Ko, K.T. Kim, Development and characterization of low-silicon cast
[2] J.Y. Song, J.C. Park, B.H. Jeong, Y.S. Ahn, Fatigue behaviour of A356 aluminium aluminum alloys for thermal dissipation, J. Alloy. Compd. 644 (2015) 673–686,
alloy for automotive wheels, Int J. Cast Met. Res. 25 (2012) 26–30, https://doi. https://doi.org/10.1016/j.jallcom.2015.04.230
org/10.1179/1743133611Y.0000000009 [13] X. Dong, H. Yang, X. Zhu, S. Ji, High strength and ductility aluminium alloy
[3] S.M.A.K. Mohammed, D.J. Li, X.Q. Zeng, D.L. Chen, Low-cycle fatigue behavior of a processed by high pressure die casting, J. Alloy. Compd. 773 (2019) 86–96,
newly developed cast aluminum alloy for automotive applications, Fatigue Fract. https://doi.org/10.1016/j.jallcom.2018.09.260
Eng. Mater. Struct. 42 (2019) 1912–1926, https://doi.org/10.1111/ffe.13035 [14] W. Jiang, Z. Fan, Y. Dai, C. Li, Effects of rare earth elements addition on micro­
[4] W. Jiang, Z. Fan, D. Liu, D. Liao, X. Dong, X. Zong, Correlation of microstructure structures, tensile properties and fractography of A357 alloy, Mater. Sci. Eng. A
with mechanical properties and fracture behavior of A356-T6 aluminum alloy 597 (2014) 237–244, https://doi.org/10.1016/j.msea.2014.01.009
fabricated by expendable pattern shell casting with vacuum and low-pressure, [15] M. Kalka, J. Adamiec, Complex procedure for the quantitative description of an
gravity casting and lost foam casting, Mater. Sci. Eng. A 560 (2013) 396–403, Al–Si cast alloy microstructure, Mater. Charact. 56 (2006) 373–378, https://doi.
https://doi.org/10.1016/j.msea.2012.09.084 org/10.1016/j.matchar.2006.01.017

10
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

[16] Y. Lu, L.A. Godlewski, J.W. Zindel, A. Lee, Use of reactive nanostructured che­ [41] C. Li, Y. Wu, H. Li, X. Liu, Microstructural formation in hypereutectic Al-Mg2Si
micals for refinement of Si eutectic in an aluminum casting alloy, J. Mater. Sci. with extra Si, J. Alloy. Compd. 477 (2009) 212–216, https://doi.org/10.1016/j.
54 (2019) 12818–12832, https://doi.org/10.1007/s10853-019-03799-9 jallcom.2008.10.061
[17] İ. Öztürk, G. Hapçı Ağaoğlu, E. Erzi, D. Dispinar, G. Orhan, Effects of strontium [42] S. Kores, M. Vončina, B. Kosec, J. Medved, Formation of AlFeSi phase in AlSi12
addition on the microstructure and corrosion behavior of A356 aluminum alloy, J. alloy with Ce addition, Metalurgija 51 (2012) 216–220.
Alloy. Compd. 763 (2018) 384–391, https://doi.org/10.1016/j.jallcom.2018.05.341 [43] M.R. Jandaghi, H. Pouraliakbar, G. Khalaj, M.-J. Khalaj, A. Heidarzadeh, Study on
[18] J. Wang, Z. Guo, J.L. Song, W.X. Hu, J.C. Li, S.M. Xiong, Morphology transition of the post-rolling direction of severely plastic deformed aluminum-manganese-
the primary silicon particles in a hypereutectic A390 alloy in high pressure die silicon alloy, Arch. Civ. Mech. Eng. 16 (2016) 876–887, https://doi.org/10.1016/j.
casting, Sci. Rep. 7 (2017) 14994, https://doi.org/10.1038/s41598-017-15223-w acme.2016.06.005
[19] G.K. Sigworth, The modification of Al–Si casting alloys: important practical and [44] Q. Wang, H. Chen, Z. Zhu, P. Qiu, Y. Cui, A characterization of microstructure and
theoretical aspects, Int. J. Metalcasting 2 (2008) 19–40, https://doi.org/10.1007/ mechanical properties of A6N01S-T5 aluminum alloy hybrid fiber laser-MIG
BF03355425 welded joint, Int J. Adv. Manuf. Technol. 86 (2016) 1375–1384, https://doi.org/10.
[20] M. Timpel, N. Wanderka, R. Schlesiger, T. Yamamoto, N. Lazarev, D. Isheim, 1007/s00170-015-8280-y
G. Schmitz, S. Matsumura, J. Banhart, The role of strontium in modifying alu­ [45] M. Bauccio, ASM Engineered Materials Reference Book, CRC, 1994.
minium-silicon alloys, Acta Mater. 60 (2012) 3920–3928, https://doi.org/10. [46] S. Amirkhanlou, S. Ji, Casting lightweight stiff aluminum alloys: a review, Crit.
1016/j.actamat.2012.03.031 Rev. Solid State Mater. Sci. 45 (2020) 171–186, https://doi.org/10.1080/10408436.
[21] M. Zamani, S. Seifeddine, Determination of optimum Sr level for eutectic Si 2018.1549975
modification in Al–Si cast alloys using thermal analysis and tensile properties, [47] F. Lasagni, H.P. Degischer, Enhanced Young’s Modulus of Al–Si alloys and re­
Int J. Metalcasting 10 (2016) 457–465, https://doi.org/10.1007/s40962-016- inforced matrices by co-continuous structures, J. Compos. Mater. 44 (2010)
0032-8 739–755, https://doi.org/10.1177/0021998309347649
[22] G. Asghar, L. Peng, P. Fu, L. Yuan, Y. Liu, Role of Mg2Si precipitates size in de­ [48] M. Sahoo, R.W. Smith, Mechanical properties of unidirectionally solidified AI-Si
termining the ductility of A357 cast alloy, Mater. Des. 186 (2020) 108280, eutectic alloys, Met. Sci. 9 (1975) 217–222, https://doi.org/10.1179/
https://doi.org/10.1016/j.matdes.2019.108280 030634575790444874
[23] G. Sha, K. O’Reilly, B. Cantor, J. Worth, R. Hamerton, Growth related metastable [49] A. Mandal, M.M. Makhlouf. Improving Aluminum Casting Alloy and Process
phase selection in a 6xxx series wrought Al alloy, Mater. Sci. Eng. A 304–306 Competitiveness. ACRC Rep 2007.
(2001) 612–616, https://doi.org/10.1016/S0921-5093(00)01545-8 [50] P. Zhang, Z. Li, B. Liu, W. Ding, Tensile properties and deformation behaviors of a
[24] E. Lee, B. Mishra, Effect of cooling rate on the mechanical properties of AA365 new aluminum alloy for high pressure die casting, J. Mater. Sci. Technol. 33
aluminum alloy heat-treated under T4, T5, and T6 conditions, Int J. Metalcasting (2017) 367–378, https://doi.org/10.1016/j.jmst.2016.02.013
12 (2018) 449–456, https://doi.org/10.1007/s40962-017-0195-y [51] X. Wu, Y. Zhu, Heterogeneous materials: a new class of materials with un­
[25] E. Lee, B. Mishra, Effect of solidification cooling rate on mechanical properties precedented mechanical properties, Mater. Res Lett. 5 (2017) 527–532, https://
and microstructure of Al–Si-Mn-Mg alloy, Mater. Trans. 58 (2017) 1624–1627, doi.org/10.1080/21663831.2017.1343208
https://doi.org/10.2320/matertrans.M2017170 [52] X. Jiang, L. Zhang, L. Zhang, T. Huang, G. Wu, X. Huang, O.V. Mishin,
[26] Z. Yuan, Z. Guo, S. Xiong, Microstructure evolution of modified die-cast Heterogeneous microstructure and enhanced mechanical properties in annealed
AlSi10MnMg alloy during solution treatment and its effect on mechanical multilayered IF steel, Mater. Sci. Eng. A 759 (2019) 262–271, https://doi.org/10.
properties, Trans. Nonferrous Met Soc. China 29 (2019) 919–930, https://doi.org/ 1016/j.msea.2019.05.034
10.1016/S1003-6326(19)65001-6 [53] A. Hazrati, Multi-scale analysis of nonlinear fatigue damage behaviour of a quad-
[27] E. Vandersluis, A. Lombardi, C. Ravindran, A. Bois-Brochu, F. Chiesa, R. MacKay, core sandwich panel with heterogeneous aluminium sheets, Theor. Appl. Fract.
Factors influencing thermal conductivity and mechanical properties in 319 Al Mech. 99 (2019) 79–94, https://doi.org/10.1016/j.tafmec.2018.11.003
alloy cylinder heads, Mater. Sci. Eng. A 648 (2015) 401–411, https://doi.org/10. [54] Y. Jiao, Y. Zhang, S. Ma, D. Sang, Y. Zhang, J. Zhao, S. Wu, Y. Liu, S. Yang, Effects of
1016/j.msea.2015.09.091 microstructural heterogeneity on fatigue properties of cast aluminum alloys, J.
[28] M. Zuo, D. Zhao, X. Teng, H. Geng, Z. Zhang, Effect of P and Sr complex mod­ Cent. South Univ. 27 (2020) 674–697, https://doi.org/10.1007/s11771-020-
ification on Si phase in hypereutectic Al-30Si alloys, Mater. Des. 47 (2013) 4323-0
857–864, https://doi.org/10.1016/j.matdes.2012.12.054 [55] Y.N. Zan, Y.T. Zhou, Z.Y. Liu, G.N. Ma, D. Wang, Q.Z. Wang, W.G. Wang, B.L. Xiao,
[29] Y.H. Cho, H.-C. Lee, K.H. Oh, A.K. Dahle, Effect of strontium and phosphorus on Z.Y. Ma, Enhancing strength and ductility synergy through heterogeneous
eutectic Al–Si nucleation and formation of β-Al5FeSi in hypoeutectic Al–Si structure design in nanoscale Al2O3 particulate reinforced Al composites, Mater.
foundry alloys, Metall. Mater. Trans. A 39 (2008) 2435–2448, https://doi.org/10. Des. 166 (2019) 107629, https://doi.org/10.1016/j.matdes.2019.107629
1007/s11661-008-9580-8 [56] J.R. Davis, et al., Aluminum and Aluminum Alloys, ASM international, 1993.
[30] F. Casarotto, A.J. Franke, R. Franke, High-pressure die-cast (HPDC) aluminium [57] Z. Hu, L. Wan, S. Wu, H. Wu, X. Liu, Microstructure and mechanical properties of
alloys for automotive applications. Adv Mater Automot Eng, 2012, 109–149, high strength die-casting Al-Mg-Si-Mn alloy, Mater. Des. 46 (2013) 451–456.
https://doi.org/10.1533/9780857095466.109 [58] M. Kutz, Handbook of Materials Selection, John Wiley & Sons, 2002.
[31] L. Hengcheng, S. Yu, S. Guoxiong, Restraining effect of strontium on the crys­ [59] H. Liao, Y. Sun, G. Sun, Correlation between mechanical properties and amount
tallization of Mg 2Si phase during solidification in Al–Si-Mg casting alloys and of dendritic α-Al phase in as-cast near-eutectic Al-11.6% Si alloys modified with
mechanisms, Mater. Sci. Eng. A 358 (2003) 164–170, https://doi.org/10.1016/ strontium, Mater. Sci. Eng. A 335 (2002) 62–66, https://doi.org/10.1016/S0921-
S0921-5093(03)00276-4 5093(01)01949-9
[32] A. Assadiki, V.A. Esin, M. Bruno, R. Martinez, Stabilizing effect of alloying elements [60] X.P. Niu, B.H. Hu, I. Pinwill, H. Li, Vacuum assisted high pressure die casting of
on metastable phases in cast aluminum alloys by CALPHAD calculations, Comput. aluminium alloys, J. Mater. Process. Technol. 105 (2000) 119–127, https://doi.
Mater. Sci. 145 (2018) 1–7, https://doi.org/10.1016/j.commatsci.2017.12.056 org/10.1016/S0924-0136(00)00545-8
[33] E. Mohammadi Mazraeshahi, B. Nami, S.M. Miresmaeili, S.M. Tabatabaei, Effect [61] C.F. Ferrarini, C. Bolfarini, C.S. Kiminami, F.W.J. Botta, Microstructure and me­
of Si on the creep properties of AZ61 cast magnesium alloy, Mater. Des. 76 (2015) chanical properties of spray deposited hypoeutectic Al–Si alloy, Mater. Sci. Eng. A
64–70, https://doi.org/10.1016/j.matdes.2015.03.021 375–377 (2004) 577–580, https://doi.org/10.1016/j.msea.2003.10.062
[34] S. Qu, A. Feng, L. Geng, J. Shen, D. Chen, Silicon nitride whisker-reinforced alu­ [62] M. Tiryakioğlu, J. Campbell, N.D. Alexopoulos, Quality indices for aluminum alloy
minum matrix composites: twinning and precipitation behavior, Metals 10 castings: a critical review, Metall. Mater. Trans. B 40 (2009) 802–811, https://doi.
(2020) 420, https://doi.org/10.3390/met10030420 org/10.1007/s11663-009-9304-5
[35] M. Shamanian, H. Mostaan, M. Safari, J.A. Szpunar, EBSD study on grain [63] A.M.A. Mohamed, A.M. Samuel, F.H. Samuel, H.W. Doty, Influence of additives on
boundary and microtexture evolutions during friction stir processing of A413 the microstructure and tensile properties of near-eutectic Al-10.8%Si cast alloy,
cast aluminum alloy, J. Mater. Eng. Perform. 25 (2016) 2824–2835, https://doi. Mater. Des. 30 (2009) 3943–3957, https://doi.org/10.1016/j.matdes.2009.05.042
org/10.1007/s11665-016-2141-1 [64] E. Czekaj, J. Zych, Z. Kwak, A. Garbacz-Klempka, Quality index of the AlSi7Mg0.3
[36] S. Chankitmunkong, D.G. Eskin, C. Limmaneevichitr, Constitutive behavior of an aluminium casting alloy depending on the heat treatment parameters, Arch.
AA4032 piston alloy with Cu and Er additions upon high-temperature com­ Foundry Eng. 16 (2016) 25–28, https://doi.org/10.1515/afe-2016-0043
pressive deformation, Met. Mater. Trans. A 51 (2020) 467–481, https://doi.org/ [65] F. Khomamizadeh, A. Ghasemi, Evaluation of quality index of A-356 aluminum
10.1007/s11661-019-05482-9 alloy by microstructural analysis, Sci. Iran. 11 (2004) 386–391.
[37] I.N.A. Oguocha, A.A. Tiamiyu, M. Rezaei, A.G. Odeshi, J.A. Szpunar, Experimental [66] E. Erzi, O. Gursoy, C. Yuksel, M. Colak, D. Dispinar, Determination of acceptable
investigation of the dynamic impact responses of as-cast and homogenized quality limit for casting of A356 aluminium alloy: supplier’s quality index (SQI),
A535 aluminum alloy, Mater. Sci. Eng. A 771 (2020) 138536, https://doi.org/10. Metals 9 (2019) 957, https://doi.org/10.3390/met9090957
1016/j.msea.2019.138536 [67] N. Afrin, D.L. Chen, X. Cao, M. Jahazi, Strain hardening behavior of a friction stir
[38] P.L. Niu, W.Y. Li, A. Vairis, D.L. Chen, Cyclic deformation behavior of friction-stir- welded magnesium alloy, Scr. Mater. 57 (2007) 1004–1007, https://doi.org/10.
welded dissimilar AA5083-to-AA2024 joints: effect of microstructure and 1016/j.scriptamat.2007.08.001
loading history, Mater. Sci. Eng. A 744 (2019) 145–153, https://doi.org/10.1016/j. [68] S.M. Chowdhury, D.L. Chen, S.D. Bhole, X. Cao, E. Powidajko, D.C. Weckman,
msea.2018.12.014 Y. Zhou, Tensile properties and strain-hardening behavior of double-sided arc
[39] M.V. Glazoff, A. Khvan, V.S. Zolotorevsky, N.A. Belov, A. Dinsdale, Casting welded and friction stir welded AZ31B magnesium alloy, Mater. Sci. Eng. A 527
Aluminum Alloys: Their Physical and Mechanical Metallurgy, Butterworth- (2010) 2951–2961, https://doi.org/10.1016/j.msea.2010.01.031
Heinemann, 2018. [69] N. Tahreen, D.L. Chen, M. Nouri, D.Y. Li, Effects of aluminum content and strain rate
[40] S. Roy, L.F. Allard, A. Rodriguez, T.R. Watkins, A. Shyam, Comparative evaluation on strain hardening behavior of cast magnesium alloys during compression, Mater.
of cast aluminum alloys for automotive cylinder heads: part I-microstructure Sci. Eng. A 594 (2014) 235–245, https://doi.org/10.1016/j.msea.2013.11.078
evolution, Metall. Mater. Trans. A 48 (2017) 2529–2542, https://doi.org/10.1007/ [70] P. Ludwik, Elemente der Technologischen Mechanik, Springer Berlin Heidelberg,
s11661-017-3985-1 Berlin, Heidelberg, 1909, https://doi.org/10.1007/978-3-662-40293-1

11
S.S. Dash, D.J. Li, X.Q. Zeng et al. Journal of Alloys and Compounds 870 (2021) 159413

[71] H. Hollomon, Tensile deformation, AIME Trans. (1945). nano-lamellar AlCoCrFeNi2.1 eutectic high entropy alloy by cryo-rolling and
[72] H.W. Swift, Plastic instability under plane stress, J. Mech. Phys. Solids 1 (1952) annealing, Sci. Rep. 8 (2018) 3276, https://doi.org/10.1038/s41598-018-21385-y
1–18, https://doi.org/10.1016/0022-5096(52)90002-1 [78] P. Shi, W. Ren, T. Zheng, Z. Ren, X. Hou, J. Peng, P. Hu, Y. Gao, Y. Zhong, P.K. Liaw,
[73] C. Ren, W. Dan, Y. Xu, W. Zhang, Effects of heterogeneous microstructures on the Enhanced strength-ductility synergy in ultrafine-grained eutectic high-entropy
strain hardening behaviors of ferrite-martensite dual phase steel, Metals 8 alloys by inheriting microstructural lamellae, Nat. Commun. 10 (2019) 489,
(2018) 824, https://doi.org/10.3390/met8100824 https://doi.org/10.1038/s41467-019-08460-2
[74] P. Sathiyamoorthi, H.S. Kim, High-entropy alloys with heterogeneous micro­ [79] T. Xiong, S. Zheng, J. Pang, X. Ma, High-strength and high-ductility AlCoCrFeNi2.1
structure: processing and mechanical properties, Prog. Mater. Sci. (2020) eutectic high-entropy alloy achieved via precipitation strengthening in a het­
100709, https://doi.org/10.1016/j.pmatsci.2020.100709 erogeneous structure, Scr. Mater. 186 (2020) 336–340, https://doi.org/10.1016/j.
[75] J.A. del Valle, F. Carreño, O.A. Ruano, Influence of texture and grain size on work scriptamat.2020.04.035
hardening and ductility in magnesium-based alloys processed by ECAP and [80] G. Purcek, O. Saray, O. Kul, Microstructural evolution and mechanical properties
rolling, Acta Mater. 54 (2006) 4247–4259, https://doi.org/10.1016/j.actamat. of severely deformed Al-12Si casting alloy by equal-channel angular extrusion,
2006.05.018 Met. Mater. Int. 16 (2010) 145–154, https://doi.org/10.1007/s12540-010-0145-1
[76] X. Wu, M. Yang, F. Yuan, G. Wu, Y. Wei, X. Huang, Y. Zhu, Heterogeneous lamella [81] Z. Ma, A.M. Samuel, H.W. Doty, F.H. Samuel, On the fractography of impact-
structure unites ultrafine-grain strength with coarse-grain ductility, Proc. Natl. tested samples of Al–Si alloys for automotive alloys, Fract. Mech. Patterns Behav.
Acad. Sci. U. S. A. 112 (2015) 14501–14505, https://doi.org/10.1073/pnas.1517193112 (2016) 27–59.
[77] T. Bhattacharjee, I.S. Wani, S. Sheikh, I.T. Clark, T. Okawa, S. Guo, [82] M.F. Hafiz, T. Kobayashi, A study on the microstructure -fracture behavior rela­
P.P. Bhattacharjee, N. Tsuji, Simultaneous strength-ductility enhancement of a tions in Al–Si casting alloys, Scr. Metall. Mater. 30 (1994) 475–480.

12

You might also like