You are on page 1of 14

Materials Today Communications 26 (2021) 102099

Contents lists available at ScienceDirect

Materials Today Communications


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

Influence of the dendritic microstructure and β-Al5FeSi phase on the wear


characteristics in a horizontally solidified Al-7Si-0.4Mg-1.2Fe alloy
Angela Vasconcelos a, Hugo Azevedo b, André Barros c, Otávio Rocha b, Mirian Motta Melo a, *
a
Federal University of Itajubá, Institute of Mechanical Engineering-UNIFEI, Itajubá, 37500-903, MG, Brazil
b
Federal Institute of Education, Science and Technology of Pará-IFPA, Belém, 66093-020, PA, Brazil
c
Department of Manufacturing and Materials Engineering, University of Campinas-UNICAMP, 13083-860 Campinas, SP, Brazil

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

Keywords: In this work an Al-7Si-0.4Mg-1.2Fe alloy (wt.%) was subjected to a horizontal solidification experiment using a
Horizontal solidification water-cooled solidification device equipped with thermocouples to obtain temperature vs. time data which in
Microstructure turn allowed solidification thermal parameters, such as growth and cooling rates (VL and TR, respectively), to be
β-Al5FeSi phase
determined. In turn, tribological behavior of samples with different secondary dendritic spacing (λ2) and
Wear feature
AlSiMgFe alloys
β-Al5FeSi platelets length (βFe) were assessed by means of dry sliding wear testing performed in a rotating fixed
ball machine, with wear volume and rate (WV and WR, respectively) being the two main investigated wear
parameters. Quantitative metallography by optical and scanning electron microscopy along with energy
dispersive X-ray spectroscopy enabled both as-cast microstructures and worn craters to be characterized. More
significant variations in wear resistance of the investigated alloy were found for ranges of λ2 and βFe that lie
within 12− 20 μm and 14–36 μm, respectively, with the coarsening of the microstructure constituted by Al-rich
primary phase dendrites surrounded by α-Al + Si + θ-Mg2Si + β-Al5FeSi eutectic structures favoring reduced WV
and WR values. For λ2 and βFe values higher than 24 and 48 μm, respectively, both WV and WR stabilize assuming
constant values that depend on the sliding distance. In addition to power type equations relating λ2 and βFe to VL
and TR, mathematical expressions for variations of WV and WR with λ2 and βFe are also proposed. Finally, a
comparative analysis with the literature is presented.

1. Introduction presumes a detailed understanding of the role of these Fe-rich phases on


the resulting properties.
Technologies to optimize the benefits of recycling of aluminum al­ Among the most common Fe-rich IMCs present in the as-cast
loys for casting and mechanical forming processes have been widely microstructure of commercial Al alloys, β-Al5FeSi phase has been
reported in the scientific community mainly due to the fact that the recognized as one of the most undesirable when a superior mechanical
energy required to recycle is significantly less than that of processes used behavior that associates high tensile strength with good ductility is
to obtain the same amount of primary alloy thus characterizing an required [5,12–14]. Commonly with the morphology of needles-like
outstanding solution for environmental appeals [1,2]. However, one of faceted plates, such microstructural phase favors stress concentrations
the greatest challenges facing recycling industries refers to the incor­ at its sharp points. In addition, since β-Al5FeSi particles are brittle and
poration of iron into aluminum scraps which in turn can lead to the hard, their presence in microstructure is detrimental to the machin­
inevitable formation of Fe-intermetallic phases during liquid-solid ability of castings. Hence, several studies have been dedicated to
transition [3–6]. Indeed, several Fe-rich intermetallic compounds continuing to better understand how nucleation of the needles-like
(IMCs) such as α-Al8Fe2Si or α-Al15(Fe,Mn)3Si2, β-Al5FeSi, Δ-Al4FeSi2, β-Al5FeSi phase and mechanisms of change in its morphology during
ω-Al7Cu2Fe and π-Al8Mg3FeSi6 can nucleate and growth during solidi­ solidification can affect the mechanical behavior [22–25]. Basak and
fication of recycled Al alloys depending not only on the alloy composi­ Babu [23], for example, have depicted that morphological change of
tion, but also on the thermal processing conditions [7–11]. Therefore, β-Al5FeSi-phase and Si could help recovering the strength and ductility
the optimized design of recycled Al alloys for a particular application in the recycled Al-Si alloys having high Fe content. According to Li et al.

* Corresponding author.
E-mail address: mirianmottamelo@unifei.edu.br (M.M. Melo).

https://doi.org/10.1016/j.mtcomm.2021.102099
Received 8 October 2020; Received in revised form 11 January 2021; Accepted 28 January 2021
Available online 4 February 2021
2352-4928/© 2021 Elsevier Ltd. All rights reserved.
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Table 1
Chemical analysis of the base A356 alloy used in this work to elaborate the investigated alloy.
Element (wt.%)

Al Si Fe Mg Cu Mn Cr Zn Ti

Balance 6.88 0.153 0.332 0.074 0.0005 0.0005 0.0010 0.0020

[13], although the Fe has a deleterious effect on both ultimate tensile However, for Fe additional contents up to 1.8 wt.%, the wear resistance
strength (UTS) and ductility, mainly due to the presence of the β-Al5FeSi was lowered by about 30, 35 and 55 % under applied loads of 20, 30 and
phase, Fe additions up to 0.8 % have increased the yield strength (YS) of 40 N respectively. Abouei et al. [17] have evaluated the solidification
Al7Si3Cu-(Mn;Fe;Sr) alloys, deducing an increase in the hardness. conditions and Mn addition on the wear behavior of Fe-rich eutectic
Studying Al-9Si-xFe alloys (x = 0.1, 0.4, 0.8 and 1.2 wt.%) processed by Al-Si piston alloy, and reported that the combined effects of high TR with
chill casting, Malavazi et al. [14] reported higher UTS and YS values for Mn addition have resulted in the formation of finer α-IMCs and, in turn,
Fe contents equal to 0.4 and 0.8 wt.% Fe, respectively, as well as a have improved the wear performance of the alloy. Saghafian et al. [19]
greater amount of β-Fe phase to 1.2 wt.% Fe. Ceschini et al. [12] state have investigated the effects of Fe addition and solidification conditions
that Fe additions up to 0.5 % in Al10Si2Cu alloy (wt.%) besides pro­ on the Al-Si eutectic alloy, and reported that the presence of Fe may give
moting the presence of the needle-like β phase, also induce a slight in­ rise to the formation of a needle-like β-Al5FeSi phase, whose fragile
crease in fatigue resistance. Additionally, they found that the presence of nature has led to a decrease in the mechanical characteristics of piston
α-Al15 (Fe,Mn)3Si2 phase, resulting from Mn addition, has influenced the alloys, including wear resistance. However, the addition of Mn/Fe in a
crack path, increasing in turn the resistance to fatigue. ratio of 1/2 has modified the needle-like β phase in alloys containing up
Although the deleterious effects on the tensile properties (UTS and to 1.2 wt.%.
elongation), the wear performance of Al-Si alloys with Fe addition has Bidmeshki et al. [20] have studied the effect of Mn addition on both
attracted the attention of researchers in recent years [15–21]. For Fe-rich IMCs and the wear behavior of Al-17.5 wt.%Si hypereutectic
instance, Ji et al. [5] have reported for Al-Mg-Si-Mn and Al-Mg-Si die alloys, and have shown that addition of 1.2 wt.%Fe to the base alloy
cast alloys that the increase in Fe concentration has reduced signifi­ increased the wear rate due to the formation of needles-like β-Al5FeSi
cantly the ductility, but accompanied by an improvement in yield stress intermetallic. On the other hand, the introduction of 0.6 wt.%Mn to the
(YS). Taghiabadi et al. [15] analyzed the wear behavior of Al-Si alloys Fe-rich alloy changed the needles-like β morphology and, as a conse­
due to the presence of the needles-like β-Al5FeSi phase, and the results quence, reduced the negative effect of Fe, but the Mn addition up to
showed that an increase in Fe from 0.15 wt.% in the base alloy to about 0.9 wt.% has not prevented the formation of needles-like β [20]. More
0.7 wt.% increased the hardness and improved wear resistance by about recently, Kaiser et al. [21], assessing the mechanical and wear behavior
10 % under applied loads of 20 and 40 N. The Fe addition up to 2.5 wt.% of hypereutectic in Al-Si-(Fe;Ni;Cr) automotive alloys, have depicted
has increased the hardness, but has decreased wear resistance. The au­ that Fe addition has improved the hardness of all the studied alloys due
thors [15] have controlled the negative influence of the β phase through to formation of hard β-Al5FeSi phase, but it has reduced YS, UTS and
combined effects of high cooling rate (TR) and chemical modification by wear resistance. Pouladvand et al. [26] have carried out investigation
Sr. In turn, Taghiabadi and Ghasemi [16], investigating the dry sliding with Al-xSi-1.2Fe(Mn) (x = 5− 13 wt.%) alloys and according to the
wear behavior of hypoeutectic Al-Si-(xFe) alloys, noted that as the results, in the Si range of 5− 9 wt.%, Fe-impurity promotes the formation
amount of Fe was increased to 0.7 wt.%, an improvement of 10 % was of finely distributed β-Al5FeSi IMCs in the matrix increasing its hardness
observed in the wear resistance under applied loads of 20, 30 and 40 N. and potential to support the tribolayer leading to the improvements in

Fig. 1. (a) Complete scheme of the solidification furnace, with details of the horizontal water-cooled device and (b) positioning of the thermocouples in the ingot.

2
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 2. (a) Schematic representation of the as-cast ingot of the investigated alloy, showing the longitudinal section for microstructural analysis and (b) techniques
applied in the measurement of λ2 and β-Al5FeSi particle length (βFe).

wear characteristics. According to the authors [26], the substitution of 2. Experimental procedure
fine β-Al5FeSi platelets by α-Fe Chinese scripts in Al-(5− 9)Si-1.2FeMn
(wt.%) alloys reduces the wear resistance, whereas the formation of star The Al7Si0.4Mg1.2Fe (wt.%) alloy investigated in this work was
or polyhedral-like α-Fe compounds in Al-(9–13)Si-1.2FeMn alloys im­ prepared from the as-cast ingot of a base alloy whose chemical
proves the wear resistance. composition furnished by the Brazilian Military Material Industry
As can be seen, although the literature presents studies on the wear (IMBEL) is shown in Table 1. It has been carried out by an optical
characteristics of Fe containing Al-based multicomponent alloys, studies emission spectroscopy, SpectroMaxx model. Fe powder of high purity
on directional solidification with AlSiMgFe alloys for recycling purposes was added to the base alloy in the stoichiometric proportion to achieve
are still scarce in the literature, especially as concerns to thermal and the desired Fe solute content (1.2 wt.%). The investigated alloy has been
microstructural analyzes, including the effects of the growth and cooling subjected to thermal analysis through a slow cooling in the crucible to
rates on the dendritic microstructure and, in turn, on the wear param­ check the liquidus and solidus temperatures. The results will be presented
eters. In addition, directional solidification works that perform β-Al5FeSi and analyzed in the Results and Discussions section.
particle analyzes are rare in the literature to evaluate the effects of Fig. 1a presents a complete scheme of the furnace used in the so­
thermal and microstructural solidification parameters on the length and lidification experiments to obtain the as-cast ingot of the investigated
distribution of β-Al5FeSi platelets IMCs in the as-cast microstructure of alloy, as well as the details of the water-cooled horizontal solidification
Al-Si casting alloys. device and the positioning of thermocouples (5, 10, 15, 20, 30, 50 70 e
An important aspect to be considered is about the solidification di­ 90 mm), type K, from the heat transfer surface (mold plate) of the hor­
rection. Upward solidification (with melt on top of the solid) is both izontal ingot mold, as seen in Fig. 2b. The data generated by reading the
thermally and solutally stable for alloys in which the rejected element at thermocouples have been processed by software OriginPro 2020 and
the solidification front induces a local liquid that is denser than the melt used to determine the transient thermal parameters during horizontal
[27]. In the case of downward vertical solidification, melt convection solidification. Two transient thermal parameters were studied herein,
arises during the process [27]. On the other hand, when the chill is which are: tip growth rate (VL) and solidification cooling rate (TR).
placed on the side of the mold (in the horizontal solidification), the solid Whereas VL represents the rate of advance of the liquidus isotherm into
grows perpendicularly to gravity and convective flow becomes even the liquid, TR expresses the first time-derivative of the thermal profiles
more complex since two-dimensional 2-D (in some cases 3-D) assump­ provided by the thermocouples (slope of the cooling curve) at the liq­
tions must be considered to formulate the inherent transport phenomena uidus temperature. Temperature-time data acquired during solidifica­
[28,29]. Under horizontal growth conditions, both thermal and tion considering the centerline corresponding to the longitudinal axis of
composition gradients may occur in the melt bath resulting in quite the casting were used to generate a plot of position from the metal-mold
complex movements by convection within the fluid, thereby leading to interface as a function of time corresponding to the liquidus front
effects that may also have a significant impact on solidification thermal passing by each thermocouple. A curve fitting technique on such
variables, solid/liquid (S/L) interface morphology, resulting structures, experimental points generates power functions of position as a function
and segregation [28,29]. of time. The first time-derivative of this function yields values for VL.
Thus, this work aims to present a study on the roles of the thermal The TR profile was determined using thermal data recorded immediately
parameters, dendritic microstructure length scale (represented by the after the passing of the liquidus front by each thermocouple, so that
secondary dendritic spacing - λ2) and β-Al5FeSi phase feature on the calculations of the slope of the cooling curve at the liquidus temperature
wear resistance in a horizontally solidified Al7Si0.4Mg1.2Fe alloy. could be performed. More details on the canting assembly applied as
well as the working principle have been described in previous studies
[28–31].

3
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 3. (a) Schematic representations of: (a) wear machine and (b) obtaining worn craters in as-cast samples from the heat transfer surface [32–38].

The casting ingot of Al7Si0.4M1.2Fe (wt.%) alloy, obtained from the area, as shown in Fig. 3b. Before testing, the samples were sanded to
horizontal solidification experiment, was sectioned longitudinally, 1200 # and polished in diluted alumina to 1 μm. The operating principle
sanded to # 600 and then immersed in regal water (HNO3 + HCl) to of the machine as well as the operational conditions to carry out the tests
reveal the macrostructure, as shown in Fig. 2a. Removal of samples for have been detailed in recent works [32–38].
microstructural analysis can be seen in Fig. 2b. The microstructure was During dry sliding testing, the rotating sphere composed of
revealed using the Keller reagent (10 mL of HF, 15 mL of HCl, 25 mL of AISI52100 microalloyed steel with hardness of 850 HV and diameter of
HNO3 and 50 mL of distilled water). 25.4 mm is in direct contact with the surface of the samples. The sliding
Measurements of λ2 and β-Al5FeSi phase length (βFe) were performed speed used (W) was 0.49 m/s (or W = 370 RPM) and the normal contact
in selected longitudinal sections of as-cast samples from the cooled base load applied to the tested part was 0.2 N. The tests were carried out
at 5, 10, 15, 20, 30, 40, 50, 60, 70, 80, 90 mm. Next, all the average under dry sliding, that is, without an abrasive solution to avoid the
values of λ2 and βFe, in turn, were correlated with VL and TR. Image presence of any interfacial element external to the surface of the sam­
processing MOTIC system and the Image J software were used to mea­ ples. In total, each as-cast sample was tested four times, due to the four
sure λ2 and βFe (~20) independent readings for each selected position. In times (t) assumed for the tests, equal to 7, 14, 21 and 28 min [37,38].
addition, microstructural characterization was carried out using a This has generated caps in the form of worn craters.
scanning electron microscope (SEM TESCAM, VEGA LMU) coupled to an After the tests of all samples were completed, the DIGILAB stereo­
energy-dispersive X-ray spectrometer (EDS, AZTec Energy X-Act, scope was used to analyze the craters and using ImageView image
Oxford). capture software the images of the worn craters were obtained. In turn,
In order to perform tribological analysis by correlating the thermal the craters diameters (D) were measured by the ImageJ software, and
and microstructural parameters with the wear parameters, such as the from these measurements, the WV and WR wear parameters were
volume and wear rate (WV and WR), micro-abrasive wear tests on as-cast calculated using Eqs. 1 and 2, respectively.
samples of the investigated alloy were carried out, as shown in Fig. 3.
π D4
The wear machine used in the experiments is a dry rotary ball type, WV = (1)
64R
widely used in recent studies [32–38], as seen in Fig. 3a. The samples for
the tests were selected in positions 5, 10, 15, 20, 30, 50, 70, and 90 mm WV
from the cooled base, with dimensions of 15 × 20 mm in cross-sectional WR = (2)
SD

Fig. 4. Thermal analysis resulting from the transient horizontal solidification of the investigated alloy: (a) thermal profiles for 8 thermocouples and (b) ther­
mal parameters.

4
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 5. Typical solidification structures for the Al7Si0.4Mg1.2Fe alloy (wt.%), resulting from the unsteady-state horizontal solidification process.

3. Results and discussion

3.1. Thermal and microstructural analysis

Fig. 4 a presents the results of the thermal analysis of the unsteady-


state horizontally solidified Al7Si0.4Mg1.2Fe (wt.%) Alloy. The thermal
data generated by the 8 thermocouples, as shown in Fig. 4a, were used to
experimentally determine the solidification thermal parameters, VL and
TR, as can be seen in Fig. 4b. From the intersection of the isotherm
represented by the liquidus temperature (TL) (Fig. 4a), a set of ordered
pairs (t, P) were obtained. This allowed generating a power type
adjustment curve, given by the expression P = 1.65(t)0.75, which was
the best possible fit in the points observed experimentally. The growth
rates were calculated using the d(1.65 t0.75)/dt derivative, resulting in
the expression VL = 2.89(P)− 0.57. The cooling rates were determined by
the dT/dt derivative at the times corresponding to the intersections of TL
with the temperature profiles generated by each thermocouple.
Evidently, the cooling water imposed high values of VL and TR close to
the heat transfer surface, which decrease along the length of the ingot
due to the formation of the solid layer during horizontal solidification, as
can be seen in Fig. 4b.
Fig. 6. Experimental laws of secondary dendritic growth as a function of
Figs. 5 shows the typical solidification structures of the investigated
thermal solidification parameters. alloy, in macro and micro-scales. A microstructural evolution along the
ingot shows that the observed dendritic microstructure is constituted by
a dendritic network. Therefore, finer dendritic microstructures are
Where D is the worn crater diameter, as seen in Fig. 3b, and R is the ball observed for as-cast samples analyzed from the heat transfer interface.
radius. The diameter was measured four times for each worn crater Fig. 6 shows the results of the interconnection of the secondary dendritic
along different radial positions. SD is the sliding distance, which has spacings (λ2) with the transient solidification thermal parameters. As
been calculated by: SD = W.t.2πR, for the four assumed test times were: expected, VL and TR have a high influence on the λ2 values, since that
207 m, 413 m, 620 m and 827 m [32–38]. lower λ2 can be observed for positions (P) in the as-cast ingot where VL
It is important to highlight that in order to avoid interference in the and TR come to be higher. Power mathematical expressions that corre­
results caused by any irregularities in the observed wear craters, du­ late λ2 as a function of VL and TR were generated, represented by general
plicates of the tests were carried out and the final diameter for each of equations given by λ2=Constant(VL)− 2/3 and λ2=Constant(TR)-1/3,
the positions and times analyzed was determined using the arithmetic respectively, which characterize λ2 dependence as a function of solidi­
mean of the results. In cases where there is a large disparity between the fication thermal parameters, as shown in Fig. 6. It is observed that the
first test and its duplicate, a triplicate was performed, and the crater exponents 1/3 and -2/3 are absolutely in agreement with the experi­
whose diameter showed the greatest deviation was discarded. mental predictions from the literature [4,39], as well as with the theo­
retical prediction of Bouchard-Kirkaldy [40] which proposed a
mathematical approach of λ2=f(VL) for binary alloys. It can be also
noted that the determined power-type equations for the growth λ2 as a
function of solidification parameters are inside the range reported in

5
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 7. (a) Pseudo binary phase diagram for the Al7Si0.4MgxFe alloy and (b) experimental cooling curve of the investigated alloy under slow cooling.

Fig. 8. Scanning electron microscopy with EDS element mapping for two as-cast samples of the investigated alloy in the following positions from the cooled base: (a)
5 mm e (b) 40 mm.

literature. intermetallic phases plus eutectic Si.


In order to observe the phase transformations that occur in the so­ Fig. 8 shows two scanning electron micrographs with element
lidification path of the Al7Si0.4Mg1.2FE alloy (wt.%), in this work, the mapping by EDS, for two as-cast samples, from the heat transfer surface.
pseudo binary phase diagram of the Al7Si0.4MgxFe alloy (wt.%) was A concomitant comparative analysis between Figs. 7 and 8 confirms that
plotted by the means of the Thermo-Calc computational thermodynamic the final microstructure of the investigated alloy consists of a primary
software, as shown in Fig. 7a, as well as the experimental cooling curve phase (Alα) and a mixture of interdendritic eutectic phases composed of
was generated by slow cooling during solidification of the investigated Alαeutectic + Si + (Mg2Si + β-Al5FeSi) IMCs. It is observed fine eutectic
alloy inside the crucible, as seen in Fig. 7b. A comparative analysis be­ β-Al5FeSi particles within the interdendritic regions surrounded by
tween the main points of phase transformation, represented by the eutectic Si particles for high VL and TR and lower λ2, as shown in Fig. 8a.
processes (1) to (4) in both figures, shows the microstructure evolution In addition, for such conditions, fibrous and spheroidal-like eutectic Si
during the cooling. The inflection points of the derivative of the exper­ has been noted. This is in agreement with the results from the literature
imental curve represent the beginning of precipitation phases new, as [28,30,41]. On the other hand, for low VL and TR and higher λ2, the
shown in Fig. 7b. It can be noted that the final microstructure of the eutectic Si undergoes from a “spheroidal + fibrous” morphology, to
investigated alloy is constituted by a primary phase (Alα) consisting of a acicular morphology, as well as needle-like β-Al5FeSi phase come to
dendritic network (see Fig. 5) and by a mixture of eutectic phases within precipitate out of interdendritic regions and over the primary dendritic
the interdendritic regions composed by the following Mg2Si + β-Al5FeSi phase, as secondary and primary eutectic reactions, respectively, as can

6
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 9. (a) and (b) Fe particle length (βFe) dependence as a function of the position in the as-cast ingot and thermal parameters (VL and TR), respectively, and (c) SEM
micrographs, showing the βFe variation.

be seen in Fig. 8b. Malavazi et al. [14] have depicted that both microstructures for three positions from the heat transfer interface. It is
increasing the Fe content and decreasing the cooling rate encouraged verified that there is no linear behavior of β-ICMs occupation on the
the platelets-like β-Al5FeSi phase to crystallize independently of silicon. microstructure, since the occupied fraction presents an inverse variation
They have also reported that the particles formed by the primary up to the 40 mm position, assuming minimum and maximum values in
eutectic reaction are bigger that those formed by the secondary eutectic positions 5 and 40 mm, respectively. From the 40 mm position, the
reaction. fraction of needles-like β phase grows again until 80 mm, as can be noted
As reported, β-Fe IMCs have a strong influence on the mechanical in Fig. 10a. Rakhmonov et al. [7] and Narayanan et al. [43] showed that
performance of Al-Si-based casting alloys, thus, knowing the formation high cooling rates also cause increase in the number density of β-Fe
of their microstructural characteristics, such as length and distribution needles. It was attributed to the displacement of the nucleation tem­
in the as-csat microstructure, it is of fundamental importance to estab­ perature of β-Fe towards lower temperatures with increasing cooling
lish the best morphological control process. In this sense, an analysis of rate, thus reducing the time available for the growth of β-Fe [7,42].
needle-like β-Al5FeSi phase of the unsteady-state horizontally solidified
Al7Si0.4Mg1.2Fe alloy (wt.%) has been carried out and the results 3.2. Wear behavior analysis
shown in Figs. 9 and 10. As a highlight of the present work, mathe­
matical expressions have been proposed which characterize of βFe`s Fig. 11a and b show the results of worn volume (WV) with the
dependence on VL, TR and λ2, as seen in Fig. 9a and b. SEM micro­ variation of the sliding distance (SD) and the worn caps (craters) for all
structures, seen in Fig. 9c for three positions in the ingot from the cooled samples tested under the assumed conditions. As expected, diameters of
base, show that fine and smaller βFe have been observed for higher VL worn craters and, consequently, wear volumes have been greater for
and TR values and smaller λ2, as found by Malavazi et al. [14] and longer test times. It is important to highlight that several scientific re­
Rakhmonov et al. [7]. searches [32,38,43–45] have used the micro-abrasion wear with
The occupation of needle-like β-Al5FeSi phase in the final micro­ rotating sphere due to the practicality of the test as well as the fact that it
structure of the investigated alloy, under the assumed solidification allows to analyze the wear mechanisms acting in the tribological system,
conditions, has been determined by the fraction of occupied area (%), as as well as possible transitions in the performance of the formed film in
shown in Fig. 10a. Fraction of occupied area values have been obtained the contact between the body and the counter-body. The sphere when
by binarization of as-cast microstructures, using Image J software [30, rotated produces an impression in the shape of a cap on the surface of the
38]. The images shown in Fig. 10b refer to the original and binarized worn sample, as can be seen in Fig. 11b. This cap or crater is evaluated

7
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 10. (a) Area fraction of occupied area by needle-like β-Al5FeSi particle in as-cast samples from the cooled base and (b) original and binarized SEM micro­
structures for three positions form the cooled base.

Fig. 11. (a) Variation of the worn volume with the sliding distance and (b) worn craters for the four assumed test times.

qualitatively and quantitatively, generating data that identify the wear is important to note that excellent determination coefficients (R2> 0.7)
resistance of the tested sample. were achieved for the mathematical equations obtained in the tests
Figs. 12 and 13 show the wear volume (WV) and rate (WR) de­ times equal to 7− 21 min, which suggests that the proposed expressions
pendences on the secondary dendritic spacings and the of the Fe phase for these times represent well the trend of the scattered experimental
length (βFe). Similar behavior of the WV and WR variations have been data. In turn, both WV and WR stabilize assuming constant values for λ2
observed, since for all test times an inverse mathematical relationship and βFe varying from 14 to 35.8 μm and 46.8–98 μm, respectively. From
has been found between (WV;WR) x λ2 and (WV;WR) x βFe for worn a concomitant analysis with Figs. 10 and 12, it is observed that finer and
samples closer to the cooled base, for λ2 and βFe values that lie within the smaller needle-like β-Al5FeSi particles are present in greater quantity in
ranges 12− 20 μm and 13.8–35.8 μm, respectively. This allowed to the microstructure, that is, in a greater fraction of occupied area (see
propose mathematical expressions that characterize the WV and WR Fig. 10a), significantly influencing the lower wear resistance observed in
variations as a function of λ2 and βFe, as shown inside Figs. 12 and 13. It samples closer to the cooled base. Even thpough the spherical/fibrous-

8
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 12. Correlation between worn volume and microstructural parameters: (a) λ2 and (b) βFe.

Fig. 13. Correlation between worn rate and microstructural parameters: (a) λ2 and (b) βFe.

like eutectic Si present in the finer microstructure, as noted in Fig. 8a, quantitative wear parameters to λ2 have also been proposed in previous
this was not able to improve the wear resistance. investigations that involved the tribological characterization of Al-based
Different wear characteristics have been reported by Azevedo et al. alloys [46,47].
[38] for the horizontally solidified Al7Si0.3Mg0.15Fe alloy, as can be The wear futures have been analyzed on worn craters surface by the
seen in Fig. 12a. It can be deduced from the comparative analysis that Figs. 13 to 16, wich show SEM micrographs with elements microanalysis
the investigated alloys with Fe contents equal to 1.2 wt.% (This work) by EDS for three solidified samples in positions from the heat transfer
and 0.15 wt.% [38] show the same wear resistance for times of 7 min surface, considering two test times, 7 and 28 min, which represent the
and 14 min, while for longer times (21 and 28 min) the lesser and greater severity conditions, indicated by points 1 and 2 in the
Al7Si0.4Mg1.2Fe alloy (wt.%) alloy presents low wear resistance (high corresponding figures, respectively. It is observed, from the worn cra­
WR), probably due to the comparatively higher amount of the ters, the presence of the wear adhesive and abrasive mechanisms acting
needle-like β IMCs and higher βFe values, as shown in Fig. 13a and b. simultaneously for the two investigated times, with predominance of the
Under these conditions, when the as-cast surface is subjected to the adhesive wear, since more severely worn regions have been observed on
micro-abrasive wear, the caused sliding deformations, the fragile and the surfaces of the craters, obviously for the longest test time (28 min),
hard Fe particles are fragmented and removed, promoting the instability as can be seen in Figs. 14b, 15 b and 16 b. Severe wear marks have been
of the tribolayer and, as a consequence, increasing the wear volume and observed on worn crater surfaces in as-cast samples with coarser mi­
rate. Pouladvand et al. [26] have depicted that the β-platelets are brittle crostructures, i.e, low TR and higher λ2, as noted in Figs. 15 and 16.
in nature and exhibit weak faceted interface with the matrix and, thus, Additionally, an oxide film layer that provides a grease effect is expected
when subjected to the sliding-induced surface strains, they are easily to form on the sample/sphere interface as a result of the increase in
fragmented and/or decohered from the matrix, especially for high Si temperature as sliding distance increases [49,50].
alloys. On the other hand, it is seen that the wear resistance of the Initially, harder particles in greater quantity, present in the as-cast
investigated alloy in this work stabilizes for λ2 and βFe values varying microstructure, such as needle-like β IMCs, are pulled out (or broken),
from 14 to 35.8 μm and 46.8–98 μm, respectively, while in the other causing severe wear on the surface, and at the same time transferred to
[38] it decreases with the λ2 variation, as seen in Figs. 12a. the test sphere. In this case, adhesive wear has been characterized by
In fact, λ2 alone cannot explain the improvement in wear resistance. high wear rates. It is important to note from the element tables, for all
However, there is a direct relationship between λ2 and the length of the cases analyzed, high Fe amounts in the regions represented by severe
Al5FeSi phase [22], as the results found in this investigation also reveal, wear (points 2). In turn, the hard particles that have welded to the
as seen in Fig. 9b. Thus, it can be suggested the use of λ2 as a key surface of the sphere and, with their rotating movement, scratches or
parameter to quantitatively represent the refinement of the micro­ parallel juices have been produced on the surfaces of the worn out
structure length scale. Experimental equations that relate the craters, thus characterizing the abrasive wear. Azevedo et al. [38] have

9
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 14. Wear characteristics and SEM/EDS analysis for as-cast samples at a position equal to 5 mm from the cooled base in the following test times: (a) 7 min and
(b) 28 min.

suggested for the as-cast samples of the horizontally solidified longer intermetallics and coarser dendrite arm spacing formed during
Al7Si0.3Mg0.15 (wt.%) alloy the occurrence of a wear mechanisms solidification of high Fe content Al alloys are directly related to a
transition from adhesive to abrasive, and they have indicated that the reduction of the tensile properties. For an as-cast 356-based alloy with
abrasive wear mode prevails for high TR and lower λ2, that is, for finer similar composition of the one studied herein, Dong et al. [51] have
microstructures. A transition of wear mechanism from adhesive to reported UTS values equal to 200 ± 11 MPa and 180 ± 4.9 MPa for λ2
abrasive has also been found by Botelho et al. [37] for horizontaly so­ values of 26.4 ± 3.6 μm and 46 ± 7.3 μm, respectively. As found in this
lidified Al3Ni1Bi (wt.%) alloy. work, for λ2 < 20 μm wear resistance decreased as microstructure
For commercial alloys it has been well known that the β-Al5FeSi became finer. With this in mind, it can be deduced that wear resistance
phase is hard and fragile compared with the primary Al-rich phase, so and tensile strength may show an opposite behavior. However, for λ2
that the presence of β particles in the microstructure is expected to lead higher than 24 μm, both wear volume and rate (WV and WR, respec­
to a general reduction of tensile properties such as ductility and ultimate tively) assume constant values for each sliding distance. Consequently,
tensile strength (UTS) [48]. It can also be deduced that the ductility may the optimized design of recycled Al alloys assumes a detailed under­
have an inverse behavior with the increase of the β phase. Besides, standing of the role of these Fe-rich phases on the resulting properties.

10
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 15. Wear characteristics and SEM/EDS analysis for as-cast samples at a position equal to 30 mm from the cooled base in the following test times: (a) 7 min and
(b) 28 min.
Fugure 16. Wear characteristics and SEM/EDS analysis for as-cast samples at a position equal to 90 mm from the cooled base in the following test times: (a) 7 min and
(b) 28 min.

Thus, in order to find an optimum combination of wear resistance and (1) The typical solidification microstructure has been formed by an
mechanical properties, we can deduce that: if λ2 > 24 μm, mechanical Al-rich primary phase, consisting of a dendritic network, and by a
strength assumes dominance since wear response is constant; but in the mixture of interdendritic eutectic phases composed of Alαeutectic +
case of λ2 values lower than this one, wear response variation becomes a eutectic Si + (Mg2Si + β-Al5FeSi) IMCs. It has been observed fine
factor that should be taken into account and the experimental expres­ eutectic β-Al5FeSi particles within the interdendritic regions
sions proposed in this study can be used as a reference. surrounded by fibrous and spheroidal-like eutectic Si particles for
high TR and lower λ2.
4. Conclusion (2) While the needle-like β-Al5FeSi IMCs length (βFe) dependence has
been characterized by inverse mathematical expressions with VL
The following major conclusions were obtained from the present and TR, the increase in βFe values with increasing λ2 was shown to
experimental investigation:

11
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Fig. 16. Wear characteristics and SEM/EDS analysis for as-cast samples at a position equal to 90mm from the cooled base in the following test times: (a) 7min and
(b) 28min.

be directly proportional, that is, smaller βFe were found for higher needle-like β-Al5FeSi IMCs for higher VL and TR values, and lower
VL and TR, and smaller λ2. βFe. On the other hand, with the decrease in VL and TR, as well as
(3) An inverse mathematical relationship has been found between for higher λ2 and βFe values, the wear resistance decreases and
(WV;WR) x λ2 and (WV;WR) x βFe for worn samples closer to the stabilizes as the wear volume and rate values reach constant
metal/mould surface, for λ2 and βFe values varying from 12 to values.
20 μm and 13.8–35.8 μm, respectively. On the other hand, con­
stant values of WV and WR have been observed for λ2 and βFe
values varying from 14 to 35.8 μm and 46.8–98 μm, respectively. Declaration of Competing Interest
(4) From the analysis of the worn crater surfaces, it has been noted
the presence of the wear adhesive and abrasive mechanisms The authors declare no conflict of interest.
acting simultaneously, with predominance of the adhesive wear.
(5) Finer as-cast microstructures have shown lower wear resistance Acknowledgements
due to the presence of the largest fraction of area occupied by
The authors acknowledge the financial support provided by IFPA -

12
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

Federal Institute of Education, Science and Technology of Pará, and [24] S.G. Shabestari, The effect of iron and manganese on the formation of intermetallic
compounds in aluminum-silicon Alloys, Mat. Sci. Eng. A 383 (2004) 289–298,
CNPq - National Council for Scientific and Technological Development
https://doi.org/10.1016/j.msea.2004.06.022.
(Grant 302846/2017-4 and 308021/2018-5), FAPEMIG - Minas Gerais [25] M. Mahta, M. Emamy, X. Cao, J. Campbell, Overview of β-Al5FeSi phase in Al-Si
Research Foundation (Grant APQ-01301-15). CAPES− Coordenação de Alloys, Mat. Sci. Res. Trends (2008) 251–271.
Aperfeiçoamento de Pessoal de Nível Superior-Brasil-Finance Code 001. [26] S. Pouladvand, R. Taghiabadi, F. Shahriyari, Investigation of the Tribological
Properties of AlxSi-1.2Fe(Mn) (x = 5-13 wt.%) Alloys, J. Mater. Eng. Perform. 27
(2018) 3323–3334, https://doi.org/10.1007/s11665-018-3420-9.
References [27] J.E. Spinelli, I.L. Ferreira, A. Garcia, Evaluation of heat transfer coefficients during
upward and downward transient directional solidification of Al–Si alloys, Struct.
[1] K. Das, W. Yin, The worldwide aluminum economy: the current state of the Multidiscipl. Optim. 31 (2006) 241–248, https://doi.org/10.1007/s00158-005-
industry, J. Miner, Met. Mater. Soc. 59 (2007) 57–63, https://doi.org/10.1007/ 0562-9.
s11837-007-0142-0. [28] C.R. Barbosa, J.O.M. Lima, G.M.H. Machado, H.A.M. Azevedo, F.S. Rocha, A.
[2] G. Gaustad, E. Olivetti, R. Kirchain, Improving aluminum recycling: a survey of S. Barros, O.F.L. Rocha, Relationship between Aluminum-Rich/Intermetallic
sorting and impurity removal technologies, Resour. Conserv. Recy. 58 (2012) phases and microhardness of a horizontally solidified AlSiMgFe alloy, Mater. Res.
79–87, https://doi.org/10.1016/j.resconrec.2011.10.010. 22 (2019) 1–12, https://doi.org/10.1590/1980-5373-mr-2018-0365.
[3] C.M. Dinnis, J.A. Taylor, A.K. Dahle, As-cast morphology of iron-intermetallics in [29] A. Barros, C. Cruz, A.P. Silva, N. Cheung, A. Garcia, O. Rocha, A. Moreira,
Al-Si foundry Alloys, Scr. Mater. 53 (2005) 955–958, https://doi.org/10.1016/j. Horizontally solidified Al-3 wt%Cu-(0.5 wt%Mg) alloys: tailoring thermal
scriptamat.2005.06.028. parameters, microstructure, microhardness, and corrosion behavior, Acta Metall.
[4] R. Chen, Y. Shi, Q. Xu, B. Liu, Effect of cooling rate on solidification parameters and Sin. 32 (2019) 695–709, https://doi.org/10.1007/s40195-018-0852-z.
microstructure of Al-7Si-0.3Mg-0.15Fe alloy, Trans. Nonferrous Met. Soc. China 24 [30] I.A. Magno, F.A. Souza, M.O. Costa, J.M. Nascimento, A.P. Silva, T.S. Costa, O.
(2014) 1645–1652, https://doi.org/10.1016/S1003-6326(14)63236-2. L. Rocha, Interconnection between the solidification and precipitation hardening
[5] S. Ji, W. Yang, F. Gao, D. Watson, Z. Fan, Effect of iron on the microstructure and processes of an AlSiCu alloy, Mater. Sci. Technol. 35 (2019) 1743–2847, https://
mechanical property of Al–Mg–Si–Mn and Al–Mg–Si diecast Alloys, Mat. Sci. Eng. doi.org/10.1080/02670836.2019.1591028.
A 564 (2013) 130–139, https://doi.org/10.1016/j.msea.2012.11.095. [31] M.L.N. Melo, C.L. Penhalber, N.A. Pereira, C.L. Pelliciari, C.A. Santos, Numerical
[6] B. Kim, S. Lee, S. Lee, H. Yasuda, Real-time radiographic observation of and experimental analysis of microstructure formation during stainless steels
solidification behavior of Al-Si-Cu casting alloys with the variation of Iron content, solidification, J. Mater. Sci. 42 (2007) 2267–2275, https://doi.org/10.1007/
Mater. Trans. 53 (2012) 374–379, https://doi.org/10.2320/matertrans.F- s10853-006-0827-8.
M2011834. [32] K.S. Cruz, E.S. Meza, F.A. Fernandes, J.M. Quaresma, L.C. Casteletti, A. Garcia,
[7] J. Rakhmonov, G. Timelli, F. Bonollo, Influence of grain refiner addition on the Dendritic arm spacing affecting mechanical properties and wear behavior of Al-Sn
precipitation of Fe-rich phases in secondary AlSi7Cu3Mg alloys, Int. J. Metalcast. and Al-Si alloys directionally solidified under unsteady-state conditions, Metall.
11 (2017) 294–304, https://doi.org/10.1007/s40962-016-0076-9. Mater. Trans. A 41 (2010) 972–984, https://doi.org/10.1007/s11661-009-0161-2.
[8] S. Ji, W. Yang, F. Gao, D. Watson, Z. Fan, Effect of iron on the microstructure and [33] E.S. Freitas, J.E. Spinelli, L.C. Casteletti, A. Garcia, Microstructure– wear behavior
mechanical property of Al-Mg-Si-Mn and Al-Mg-Si diecast Alloys, Mat. Sci. Eng. A correlation on a directionally solidified Al–In monotectic alloy, Tribol. Int. 66
564 (2013) 130–139, https://doi.org/10.1016/j.msea.2012.11.095. (2013) 182–186, https://doi.org/10.1016/j.triboint.2013.05.009.
[9] A.M. Samuel, H.W. Doty, S. Valtierra, F.H. Samuel, Beta Al5FeSi phase platelets- [34] E.S. Freitas, A.P. Silva, J.E. Spinelli, L.C. Casteletti, A. Garcia, Inter-relation of
porosity formation relationship in A319.2 type alloys, Int. J. Metalcast. 12 (2018) microstructural features and dry sliding wear behavior of monotectic Al–Bi and
55–70, https://doi.org/10.1007/s40962-017-0136-9. Al–Pb alloys, Tribol. Lett. 55 (2014) 111–120, https://doi.org/10.1007/s11249-
[10] A.M. Samuel, F.H. Samuel, H.W. Doty, Observations on the Formation of β-Al5FeSi 014-0338-8.
phase in 319 type Al-Si alloys, J. Mater. Sci. 31 (1996) 5529–5539, https://doi. [35] T.A. Costa, M. Dias, E.S. Freitas, L.C. Casteletti, A. Garcia, The effect of
org/10.1007/BF01159327. microstructure length scale on dry sliding wear behaviour of monotectic Al-Bi-Sn
[11] W. Khalifa, F.H. Samuel, J.E. Gruzleski, Iron intermetallic phases in the Al corner of alloys, J. Alloys. Compd. 689 (2016) 767–776, https://doi.org/10.1016/j.
the Al-Si-Fe system, Metall. Mater. Trans. A 34 (2003) 807–825, https://doi.org/ jallcom.2016.08.051.
10.1007/s11661-003-0116-y. [36] R.V. Reyes, V.E. Pinotti, C.R. Afonso, L.C. Casteletti, A. Garcia, J.E. Spinelli,
[12] L. Ceschini, I. Boromei, A. Morri, S. Seifeddine, I.L. Svensson, Effect of Fe content Processing, As-Cast microstructure and wear characteristics of a monotectic Al-Bi-
and microstructural features on the tensile and fatigue properties of the Al-Si10- Cu alloy, J. Mater. Eng. Perform. 28 (2019) 1201–1212, https://doi.org/10.1007/
Cu2 alloy, Mater. Des. 36 (2012) 522–528, https://doi.org/10.1016/j. s11665-018-3851-3.
matdes.2011.11.047. [37] T.M. Botelho, H.M. Azevedo, G.H. Machado, C.R. Barbosa, F.S. Rocha, T.A. Costa,
[13] Z. Li, N. Limodin, A. Tandjaoui, P. Quaegebeur, P. Osmond, D. Balloy, Influence of O.L. Rocha, Effect of solidification process parameters on dry sliding wear behavior
Sr, Fe and Mn content and casting process on the microstructures and mechanical of AlNiBi alloy, Trans. Nonferrous Met. Soc. China 30 (2020) 582–594, https://doi.
properties of AlSi7Cu3 alloy, Mat. Sci. Eng. A 689 (2017) 286–297, https://doi. org/10.1016/S1003-6326(20)65237-2.
org/10.1016/j.msea.2017.02.041. [38] H.M. Azevedo, T.M. Botelho, C.R. Barbosa, A.P. Sousa, T.A. Costa, O.L. Rocha,
[14] J. Malavazi, R. Baldan, A.A. Couto, Microstructure and mechanical behaviour of Study of dry wear behavior and resistance in samples of a horizontally solidified
Al9Si alloy with different Fe contents, Mater. Sci. Technol. (2014) 1743–2847, and T6/Heat‑treated automotive AlSiMg alloy, Tribol. Lett. 68 (2020), https://doi.
https://doi.org/10.1179/1743284714Y.0000000659. org/10.1007/s11249-020-01302-z.
[15] R. Taghiabadi, H.M. Ghasemi, S.G. Shabestari, Effect of iron-rich intermetallics on [39] J.O. Lima, C.R. Barbosa, I.A.B. Magno, J.M. Nascimento, A.S. Barros, M.C. Oliveira,
the sliding wear behavior of Al–Si alloys, Mat. Sci. Eng. A 490 (2008) 162–170, F.A. Souza, O.L. Rocha, Microstructural evolution during unsteady-state horizontal
https://doi.org/10.1016/j.msea.2008.01.001. solidification of Al-Si-Mg (356) alloy, Trans. Nonferrous Met. Soc. Chin. 28 (2018)
[16] R. Taghiabadi, H.M. Ghasemi, Dry sliding wear behaviour of hypoeutectic Al– Si 1073–1083, https://doi.org/10.1016/S1003-6326(18)64751-X.
alloys containing excess iron, Mater. Sci. Technol. 25 (2009) 1017–1022, https:// [40] D. Bouchard, J.S. Kirkaldy, Prediction of dendrite arm spacings in unsteady and
doi.org/10.1179/174328408X302468. steady-state heat flow of unidirectionally solidified binary alloys, Metall. Mater.
[17] V. Abouei, H. Saghafian, S.G. Shabestari, M. Zarghami, Effect of Fe-rich Trans. B 28 (1997) 651–663, https://doi.org/10.1007/s11663-997-0039-x.
intermetallics on the wear behavior of eutectic Al− Si piston alloy (LM13), Mater. [41] R. Chen, Q. Xu, H. Guo, Z. Xia, Q. Wu, B. Liu, Correlation of solidification
Des. 31 (2010) 3518–3524, https://doi.org/10.1016/j.matdes.2010.02.015. microstructure refining scale, Mg composition and heat treatment conditions with
[18] C. Lin, S. Wu, S. Lü, J. Zeng, P. An, Dry sliding wear behavior of rheocast mechanical properties in Al-7Si-Mg cast aluminum alloys, Mater. Sci. Eng. A 685
hypereutectic Al− Si alloys with different Fe contents, Trans. Nonferrous Met. Soc. (2017) 391–402, https://doi.org/10.1016/j.msea.2016.12.051.
China 26 (2016) 665–675, https://doi.org/10.1016/S1003-6326(16)64156-0. [42] L.A. Narayanan, F.H. Samuel, J.E. Gruzleski, Crystallization behavior of iron-
[19] H. Saghafian, S.G. Shabestari, S. Ghadami, M.H. Ghoncheh, Effects of Iron, containing intermetallic compounds in 319 aluminum-alloy, Metall. Mater. Trans.
manganese, and cooling rate on microstructure and dry sliding wear behavior of A 25 (1994) 1761–1773, https://doi.org/10.1007/BF02668540.
LM13 aluminum alloy, Tribol. Trans. 60 (2017) 888–890, https://doi.org/ [43] R.C. Cozza, D.K. Tanaka, R.M. Souza, Friction coefficient and abrasive wear modes
10.1080/10402004.2016.1227513. in ball-cratering tests conducted at constant normal force and constant pressure -
[20] C. Bidmeshki, V. Abouei, H. Saghafian, S.G. Shabestari, M.T. Noghani, Effect of Mn preliminary results, Wear 267 (2009) 61–70, https://doi.org/10.1016/j.
addition on Fe-rich intermetallics morphology and dry sliding wear investigation of wear.2009.01.055.
hypereutectic Al-17.5%Si alloys, J. Mater. Res. Technol 5 (2016) 250–258, https:// [44] Y.H. Cheng, T. Browne, B. Heckerman, Mechanical and tribological properties of
doi.org/10.1016/j.jmrt.2015.11.008. CrN coatings deposited by large area filtered cathodic arc, Wear 271 (2011)
[21] M.S. Kaiser, S.H. Sabbir, M.S. Kabir, M.R. Soummo, M.A. Nur, Study of mechanical 775–782, https://doi.org/10.1016/j.wear.2011.03.011.
and wear behaviour of hyper-eutectic Al-Si automotive alloy through Fe, Ni and Cr [45] R.V. Camerini, R.B. Souza, F. Carli, A.S. Pereira, N.M. Balzaretti, Ball cratering test
addition, Mater. Res. 21 (2018) 1–9, https://doi.org/10.1590/1980-5373-mr- on ductile materials, Wear 271 (2011) 770–774, https://doi.org/10.1016/j.
2017-1096. wear.2011.03.013.
[22] X. Cao, J. Campbell, Morphology of β-Al5FeSi phase in Al-Si cast alloys, Mater. [46] B.P. Reis, M. M.Lopes, A. Garcia, C.A. Santos, The correlation of microstructure
Trans. 47 (2006) 1303–1312, https://doi.org/10.2320/matertrans.47.1303. features, dry sliding wear behavior, hardness and tensile properties of Al-2wt% Mg-
[23] C.B. Basak, N.H. Babu, Morphological changes and segregation of β-Al9Fe2Si2 Zn alloys, J. Alloys. Compd. 764 (2018) 267–278, https://doi.org/10.1016/j.
phase: a perspective from better recyclability of cast Al-Si Alloys, Mater. Des. 108 jallcom.2018.06.075.
(2016) 277–288, https://doi.org/10.1016/j.matdes.2016.06.096. [47] C. Ache, M. Lopes, B. Reis, A. Garcia, C. Santos, Dendritic Spacing/Columnar grain
diameter of Al–2Mg–Zn alloys affecting hardness, tensile properties, and dry

13
A. Vasconcelos et al. Materials Today Communications 26 (2021) 102099

sliding wear in the As Cast/Heat treated conditions, Adv. Eng. Mater. (2020), [50] I. Sağlam, D. Özyürek, K. Çetinkaya, Effect of ageing treatment on wear properties
https://doi.org/10.1002/adem.201901145. and electrical conductivity of Cu-Cr-Zr alloy, Bull. Mater. Sci. 34 (2011)
[48] S. Seifeddine, S. Johansson, I.L. Svensson, The influence of cooling rate and 1465–1470, https://doi.org/10.1007/s12034-011-0344-5.
manganese content on the β-Al5FeSi phase formation and mechanical properties of [51] J.X. Dong, P.A. Karnezis, G. Durrant, B. Cantor, The effect of Sr and Fe additions on
Al–Si-based alloys, Mater. Sci. Eng. A 490 (1-2) (2008) 385–390, https://doi.org/ the microstructure and mechanical properties of a direct squeeze cast Al-7Si-0.3
10.1016/j.msea.2008.01.056. Mg alloy, Metall. Mater. Trans. A 30 (1999) 1341–1356, https://doi.org/10.1007/
[49] İ. Şimşek, D. Özyürek, Investigation of the effects of Mg amount on microstructure s11661-999-0283-6.
and wear behavior of Al-Si-Mg alloys, Eng. Sci. Technol. an Int. J. 22 (2019)
370–375, https://doi.org/10.1016/j.jestch.2018.08.016.

14

You might also like