Sunteți pe pagina 1din 7

Materials Science & Engineering A 733 (2018) 9–15

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Microstructure and mechanical properties of high-pressure-assisted T


solidification of in situ Al–Mg2Si composites
Xian Tonga,b, Dechuang Zhanga, Kun Wangb, Jianguo Lina, Yeye Liua, Zimu Shib, Yuncang Lic,
⁎ ⁎⁎
Jixing Linb,d, , Cuie Wenc,
a
School of Materials Science and Engineering, Xiangtan University, Xiangtan 411105, China
b
Department of Material Engineering, Zhejiang Industry & Trade Vocational College, Wenzhou 325003, China
c
School of Engineering, RMIT University, Melbourne, Victoria 3001, Australia
d
School of Physics and Optoelectronics Xiangtan University, Xiangtan 411105, China

A R T I C LE I N FO A B S T R A C T

Keywords: In this study, an Al–Mg2Si composite was solidified under a high pressure of 3 GPa and the effects on micro-
Al–Mg2Si composites structural evolution, mechanical properties including hardness, friction, and wear behavior, and compression
Friction and wear behavior properties were investigated. The results indicate that the composite solidified under high pressure exhibited
High-pressure solidification significantly enhanced hardness, compressive properties, wear, and friction resistance compared to the as-cast
Mechanical properties
and solid solution–treated (ST) composites. The hardness of the composites under high-pressure solidification
Microstructure
was 2.25 times and 1.37 times those of the as-cast and ST composites, respectively. The compressive yield
strength, elastic modulus, and compressive strain were 121.5 MPa, 29.6 GPa, and 22.3% for the as-cast com-
posite, 146.1 MPa, 44.5 GPa, and 35.8% for the ST composite, and 151.3 MPa, 49.3 GPa, and > 75% for the
high-pressure-solidified composite, respectively. The friction coefficient of the composite under high-pressure
solidification was only 44% of that of the as-cast composite and 67% of that of the ST composite. The wear
modes were adhesive wear and abrasive wear for the as-cast composite, but only abrasive wear for the composite
solidified under high pressure; the fracture mode transformed from cleavage fractures in the as-cast composite to
quasi-cleavage fractures in the composite solidified under high pressure.

1. Introduction [5–7], centrifugal milling [8], ultrasonic refining [9,10], semi-solid


treatment [11], and hot deformation [12].
In situ Al–Mg2Si composites are promising lightweight structural Pressure and temperature are the two crucial parameters that affect
materials for applications in abrasion-resistant part applications such as the solidification behavior of metals and alloys. In recent years, high-
aerospace, vehicle engineering, and electronics packaging due to their pressure technology has been widely investigated as an auxiliary in the
unique combination of excellent physical and mechanical properties manufacturing of materials with improved performance. In particular,
including low density, high specific strength and wear resistance, and not only can the grain size of an alloy be effectively reduced, but also
excellent thermal stability [1,2]. However, conventionally solidified the phase structure and composition can be adjusted during solidifi-
Al–Mg2Si composites exhibit coarse and irregular-shaped primary cation under higher pressure because a non-equilibrium solidified mi-
Mg2Si phases, leading to stress concentration at the interface between crostructure can be obtained which is difficult to obtain via conven-
the primary Mg2Si phase and the α-Al matrix, which results in low tional solidification. Alloys solidified under high pressure may exhibit
mechanical strength and toughness. This drawback has so far limited high density with reduced concentrations of defects, thus possessing
the industrial applications of Al–Mg2Si composites. In recent years, enhanced material properties [13]. This technology has been used in
many methods have been attempted in order to modify the size, mor- the solidification of some alloys in the past, such as Al–Si [14], Al–Mg
phology, and distribution of the primary Mg2Si phase to improve the [15], Al–Ge [16], Al–Cu [17], Mg–Zn–Y [18], and Ti–Al matrix alloys
overall performance of the composite, including adding modifiers or [19]. However, high-pressure-assisted solidification of Al–Mg2Si com-
grain refiners during casting [3,4], solid-solution and aging treatment posites has rarely been investigated to date. In this study, the influence


Corresponding author at: Department of Material Engineering, Zhejiang Industry & Trade Vocational College, Wenzhou 325003, China
⁎⁎
Corresponding author.
E-mail addresses: linjixing@163.com (J. Lin), cuie.wen@rmit.edu.au (C. Wen).

https://doi.org/10.1016/j.msea.2018.07.032
Received 31 May 2018; Received in revised form 4 July 2018; Accepted 10 July 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

Table 1
Chemical compositions of Al-20% Mg2Si composite (wt%).
Composites Mg Si P Fe Impurity Al

Al-20% Mg2Si 12.89 7.58 0.003 0.17 0.04 Balance.

of high-pressure solidification on the microstructural evolution and


mechanical properties of an in situ Al–Mg2Si composite has been in-
vestigated. Our results indicate that high-pressure-assisted solidifica-
tion is a promising technique in achieving excellent wear and friction
performance, high mechanical strength, and extraordinarily high
plastic deformation ability in Al–Mg2Si composites.

2. Experimental procedure

The in situ Al–20% Mg2Si (wt%) composite was prepared by melting


pure Al (99.7 wt%), high-purity Mg (99.93 wt%), pure Si (99.5 wt%),
and Al-3.5 P intermediate alloy (Shandong Shanda Al & Mg Melt
Technology Co. Ltd, China). A graphite crucible in a high-purity argon
atmosphere was heated to 780 °C for the melting. The molten alloy was
cooled below 720 °C, then poured into a 250 °C preheated cast-steel
mold with an inner diameter of 70 mm and length of 80 mm. The
chemical composition of the Al–Mg2Si composite ingot was measured
using a wavelength dispersive X-ray fluorescence spectrophotometer
(XRF, S4 Pioneer, Bruker, Germany) and is listed in Table 1. High-
pressure-assisted solidification was performed using a six-anvil cubic
high-pressure apparatus (CS-1B, Zhengzhou Research Institute for
Abrasive & Grinding Co. Ltd, China). Cylindrical samples of the com-
posite with a diameter of 10 mm and length of 25 mm ground with
1000 grit silicon carbide sandpaper were prepared for the high-pres-
sure-assisted solidification. The composite cylinders were heated to
800 °C, then a high pressure of 3 GPa was applied for 20 min, followed
by rapid cooling to room temperature. Conventional ST was carried out
at 540 °C for 6 h, followed by quenching in water at room temperature.
The microstructure of the composite samples was observed by op-
tical microscopy (OM, DM2500C, Leica, Germany) after etching with
5% HF solution (5 ml HF, 95 ml distilled water, in vol%). X-ray dif-
fraction analysis (XRD, D/max 2500, Rigaka, Japan) was used to
characterize the phase constituents with Cu-kα radiation
(λ = 1.5406 nm), executed in the 2θ range from 10° to 90° with a
scanning speed of 4 min-1 at room temperature. The morphology, sur-
face roughness, and chemical composition of local regions of the
composite samples were analyzed by scanning electron microscopy
(SEM, Pro X FEI, Phenom, Netherlands) and energy dispersive spectrum
analysis (EDS, X-Max, Oxford, England) at 15 kV. The thermal proper-
ties of the composite samples were determined using differential
scanning calorimetry (DSC, SDT-Q650 TA, USA) in a high-purity argon Fig. 1. Microstructure of Al-20% Mg2Si composite under different solidification
atmosphere at a heating speed of 10 °C/min from room temperature to conditions of: (a) as-cast, (b) as-cast followed by solution treatment (ST), (c)
high pressure of 3 GPa assisted.
800 °C. The hardness of the composite samples was measured using a
Vickers micro-hardness tester (MicroMet 6000, Buehler, USA) with a
testing load of 200 g and a dwelling time of 15 s. The friction and wear 3. Results and discussion
behavior of the composite samples was investigated using high-speed-
reciprocating friction equipment (HSR-2M Lanzhou Zhongke Kaihua 3.1. Microstructural evolution of Al–Mg2Si composites under high-pressure-
Technology Development Co. Ltd, China) at room temperature with a assisted solidification
load of 5 N, a stroke of 4 mm, a wear rate of 1 Hz, and a testing time of
15 min. The sample surface was polished to a surface roughness of The microstructures of the Al–20% Mg2Si composites under dif-
0.3 µm (Ra) before wear testing. The weights of the composite samples ferent solidification conditions are shown in Fig. 1. It can be seen that
before and after friction and wear testing were determined by an the black primary Mg2Si phase and the gray eutectic Mg2Si phase were
electronic scale with accuracy to 0.1 mg. The compression properties of uniformly distributed in the white α-Al matrix in the as-cast composite
the composite samples were tested using a universal testing machine (Fig. 1a). The primary Mg2Si phase exhibited an irregular polygonal
(Instron 3369, USA) at a deformation rate of 0.5 mm/min at room shape with sharp corners and its average size reached about
temperature. The compression samples were machined to a diameter of 37.2 ± 7.1 µm; while the eutectic Mg2Si phase showed a fibrous
3 mm and length of 6 mm. The surface morphology of the compression morphology. After ST under atmospheric pressure, the primary Mg2Si
fractures and wear was imaged by SEM. phase became rounder and the average size of the phase reduced to

10
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

35.3 ± 8.2 µm; while the eutectic Mg2Si phase transformed into a fine
granular and short rod-like morphology (Fig. 1b). It can be deduced
that rapid atomic migration occurred in the composite during ST be-
cause there was a concentration gradient from high to low from the tip
to the flat regions of the primary Mg2Si phase. Atomic diffusion from
the high-concentration region at the tip to the flat region with low
concentration caused the sharp corners of most of the primary Mg2Si
phase to transform into round and blunt phase morphologies. The mi-
crostructure of the Al–Mg2Si composite changed significantly under the
high pressure of 3 GPa–assisted solidification, compared to those of the
solidified composite under atmospheric pressure and the ST composite.
It can be seen that the primary Mg2Si phase then disappeared and an
elliptical primary α-Al phase was precipitated with an average size of
approximately 59.3 ± 10.2 µm (Fig. 1c). A large number of fine eu-
tectic Mg2Si phase particles were distributed around the α-Al phase in
the form of a reticulation; while some of the eutectic Mg2Si phase was
distributed within the α-Al in the form of granules. The eutectic Mg2Si Fig. 3. DSC curves of Al-Mg2Si composites prepared via conventional casting
phase of the composites after high-pressure-assisted solidification ex- and 3 GPa high pressure assisted solidification.
hibited a denser distribution and finer sizes compared to those of the as-
cast and ST composites. Al–20 wt% Si alloy.
The XRD patterns of the Al–20Mg2Si composites under different The DSC curves of the as-cast and 3 GPa high-pressure-solidified
solidification conditions are shown in Fig. 2. As can be seen, the com- Al–Mg2Si composites are shown in Fig. 3. As can be seen, the initial
posites in the conventional as-cast and ST states were mainly composed temperature of the exothermic peak and the peak temperature were
of α-Al and Mg2Si phases. After ST, the peak intensity of (311) dif- 614.2 °C and 603.9 °C for the as-cast composite sample, and 622.6 °C
fraction of the α-Al phase increased, while the intensities of the other and 611.6 °C for the composite sample after high-pressure solidifica-
peaks of the α-Al phase decreased or even disappeared. However, after tion, respectively. These temperatures correspond to the eutectic
high-pressure solidification at 3 GPa, almost no Mg2Si phase was de- starting point and the eutectic reaction of the eutectic Mg2Si phase,
tected in the XRD pattern of the composite, while the peak intensities of which indicates that the eutectic Mg2Si phase of the composite after
the α-Al phase were significantly increased and their diffraction angle high-pressure solidification exhibited a higher nucleation temperature
shifted right to a higher angle, as can be seen clearly in the magnified and a reduced nucleation rate, leading to a greater degree of super-
angular range. cooling. This resulted in the finer grain size of the eutectic phase. The
According to the Bragg diffraction equation [20]: increase in the eutectic point temperature caused a deviation in the
2 d Sin θ = n λ (1) eutectic point composition of the composite during solidification,
leading to the occurring of the so-called divorced eutectic phenomenon
where λ is the wavelength of the X-ray (a constant), θ is the scattering [21]. The primary α-Al phase experienced a greater degree of super-
angle, n is an integer representing the order of the diffraction peak, and cooling and a faster growth rate relative to the eutectic Mg2Si phase.
d is the interplanar spacing generating the diffraction. Thus, the primary α-Al phase nucleated and grew preferentially during
It can be seen that the interplanar spacing (d) decreases with an the solidification process. This explains why the eutectic Mg2Si phase
increase in the diffraction angle. In this study, the reason for the right exhibited a more refined microstructure in the composite after high-
shift of the diffraction angle is twofold. On the one hand, Si has a pressure solidification. Regarding the effect of high pressure on the
smaller lattice constant (1.11 Å) than Al (1.43 Å), so it became a solid occurrence of the primary phase (that is, whether it was a primary α-Al
solution in the Al matrix under high pressure and replaced the Al, re- phase or a primary Mg2Si phase) and its morphology, the Clausiu-
sulting in a decrease in the lattice constant of the composite, which led s–Clapeyron law can be used to estimate the heat of phase transition
to a decrease in the interplanar spacing; thus the diffraction peak from the vapor pressures measured at two temperatures and therefore
shifted to a higher angle. Ma et al. [14] observed a similar result of a to determine the occurrence of the primary phase [22]. Jie et al. [23]
right shift of α-Al diffraction peaks in a high-pressure-solidified reported that the melting point of a primary α-Al phase increased with
increasing pressure at a ratio of ~ 64 °C per GPa. When the solidifica-
tion pressure reaches a certain value, the melting point of Al will be
higher than that of the primary Mg2Si phase. At this point, the α-Al
phase will be preferentially precipitated as the primary phase during
solidification of the composite.
The EDS analyses of Si and Mg elements in the α-Al solid solution of
Al–20Mg2Si composites under different solidification conditions are
shown in Fig. 4. As can be seen, the average content of Mg and Si in the
α-Al matrix was measured at 2.85 and 0.61 wt% for the as-cast alloy,
respectively (Fig. 4a). After ST, the average content of Mg and Si
reached 3.01 and 1.28 wt%, respectively (Fig. 4b). According to the
pseudo-binary phase diagram of Al–Mg2Si [24], the solid solubility of
an Mg2Si phase in an α-Al matrix is about 1.6 wt% after solid solution
at 540 °C under atmospheric pressure; the content of Mg and Si at this
temperature was 3.01 and 1.28 wt%, similar to the values indicated in
the pseudo-binary phase diagram. When the composite was solidified at
a pressure of 3 GPa, the average content of Mg and Si increased to
6.90 wt% and 1.68 wt%, respectively, which is 142% and 175% higher
Fig. 2. XRD patterns of the Al-20Mg2Si composites under different solidifica-
than in the as-cast condition (Fig. 4c). The above results indicate that
tion conditions of: (a) as-cast, (b) ST, (c) high pressure of 3 GPa assisted.

11
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

Fig. 4. EDS analysis of Si and Mg elements in the α-Al solid solution of Al-20Mg2Si composites under different solidification conditions: (a) as-cast, (b) ST, (c) high
pressure of 3 GPa assisted.

the solid solubility of Mg and Si in the α-Al matrix after high-pressure the Al–Mg2Si composites under different solidification conditions are
solidification increased to a much higher level than under a normal shown in Fig. 5. It can be seen that the friction coefficient of the as-cast
pressure. Under the high-pressure condition, the diffusion coefficient of and ST composites fluctuated in the initial stage and then gradually
solutes such as Mg and Si in the α-Al matrix was reduced and the so- tended toward a steady state, while the composite samples after high-
lidification rate was accelerated; therefore the solid solubility of the pressure-assisted solidification maintained stability. The average fric-
solute elements in the matrix was increased [25]. Since the atomic tion coefficient and wear loss trends of the composites are in the order:
radius of Si (1.11 Å) is smaller than that of Mg (1.60 Å), Si is more easily as-cast > ST > high-pressure solidification. The friction coefficient va-
dissolved into the α-Al matrix. Therefore, the increase in the Si con- lues were 1.159, 0.754, and 0.507 for the as-cast, ST, and high-pressure-
centration in the α-Al solid solution is greater than that of Mg. This solidified composites, respectively (Fig. 5a). It can be seen that the
result is consistent with the peak position shifting in the XRD diffraction friction coefficient of the composite samples under high-pressure soli-
patterns (Fig. 2). dification was only 44% of that of the as-cast samples and 67% of that
of the ST samples. The wear loss values were 27.1 mg for the as-cast
3.2. Mechanical properties of Al–Mg2Si composites solidified under composite samples, 21.3 mg for the ST samples, and 15.9 mg for the
different conditions high-pressure-solidified composites. Overall, the friction coefficient and
wear loss results indicate that the wear resistance of the composites was
The friction coefficient curves, hardness, and surface roughness of improved significantly after high-pressure-assisted solidification. The

12
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

Fig. 5. Friction and wear properties of Al-Mg2Si composites under different solidification conditions: (a) friction coefficient curves, (b) hardness, friction coefficient
and surface roughness.

hardness and surface roughness values of the composites solidified particles, and became more resistant to falling off during the wear
under different conditions are shown in Fig. 5b. It can be seen that the process, thus also improving the friction resistance of the composites.
hardness values of the composites were 72.2 ± 2.0 (HV), 118.9 ± 2.7 Liu and Zhao [26] reported similar results in a study of an Al–Si alloy.
(HV), and 162.7 ± 1.5 (HV), and the surface roughness values were The composite after high-pressure solidification exhibited sig-
1.533, 1.487, and 1.473 µm for the as-cast, ST, and high-pressure-soli- nificantly increased hardness. Further, the majority of the primary
dified samples, respectively. The hardness of the composites under Mg2Si phase dissolved into the Al matrix, with a small amount homo-
high-pressure solidification was 2.25 times and 1.37 times those of the geneously embedded in the matrix with finer sizes and spheroidized
as-cast and ST composites, respectively. shapes. The spheroidized primary Mg2Si particles became more re-
The SEM images of the surface morphologies of the Al–Mg2Si sistant to falling off and thus contributed to improving the abrasion
composites after friction and wear tests are shown in Fig. 6. The as-cast resistance in the wear process. When the solid solubility of the alloy
composite sample exhibited a rough surface topography with embedded increased, it caused lattice distortion and formed a strong stress field.
primary Mg2Si particles (as shown in Fig. 6a and magnified in the in- This hindered dislocation movement during deformation and resulted
sert). The furrows on the surface of the as-cast sample were deep and in increased hardness of the composite, and this increase in hardness
wide, and decorated with fine cracks and exfoliation pits. The surface was more pronounced at a higher solid solubility as in the composite
morphology of the ST composite sample was similar to that of the as- after high-pressure solidification. In addition, after high-pressure soli-
cast composite sample, but the furrows were shallower with no obvious dification the composite became denser due to the reduced number of
cracks. In addition, there were more stripping pits and the abrasive casting defects, which also contributed to the increase in hardness of
debris in the pits was mainly composed of small particle agglomerates. the composite [27].
The wear modes of the as-cast and ST composites mainly comprised The compressive stress–strain curves of the Al–Mg2Si composites
adhesive and abrasive wear. After high-pressure solidification, the wear under different solidification conditions are shown in Fig. 7. The
surface of the composite exhibited significantly narrower and shallower compressive yield strength (σYS), compression elastic modulus (E), and
furrows with slight spalling pits. The main wear mode was abrasive compressive strain (δ) were 121.5 ± 2.7 MPa, 29.6 ± 4.6 GPa, and
wear. After ST and high-pressure-assisted solidification, on the one 22.3 ± 1.3% for the as-cast composite, 146.1 ± 2.9 MPa,
hand, the hardness of the composites increased and their greater 44.5 ± 3.7 GPa, and 35.8 ± 1.6% for the ST composite, and
hardness provided excellent ability to resist plastic deformation, thus 151.3 ± 1.8 MPa, 49.3 ± 3.4 GPa, and > 75% for the composite after
possessing enhanced wear resistance; on the other hand, the eutectic high-pressure solidification. It is worth noting that the Al–Mg2Si com-
Mg2Si phase formed short rods and the primary Mg2Si phase corners posite after high-pressure solidification showed no cracks or fractures
became round, leading to a weakened stress concentration effect at the during compression.
interface between the primary phase particles and the Al matrix. The fractography of the Al–Mg2Si composites after compression
Therefore, the initiation and propagation of cracks were slowed down tests are shown in Fig. 8. The as-cast and ST composite samples showed
during the wear process. The round primary Mg2Si phase also enhanced a shear failure mode with brittle fracture characteristics and an azimuth
the bonding strength between the Al matrix and the reinforcing of ~ 45° between the fracture plane and the compression axis.

Fig. 6. SEM images of surface morphologies of Al-Mg2Si composites under different solidification conditions: (a) as-cast, (b) ST, (c) high pressure of 3 GPa assisted.

13
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

stress of a material is proportional to the square root of the solute


element concentration in solid solution. When the solid solubility of an
alloy increases, the lattice distortion increases and the dislocation
movement is hindered, leading to the enhanced yield strength of the
alloy [29]. In addition, the corner blunting of the primary Mg2Si phase
and the refinement of the eutectic Mg2Si phase reduced the stress
concentration during the deformation of the composites after ST, re-
sulting in increases in strength and plastic deformation ability due to
the increased bonding strength with the matrix. After high-pressure-
assisted solidification, the primary Mg2Si phase disappeared, the eu-
tectic Mg2Si phase was refined, and the α-Al phase became elliptical in
shape, resulting in improved plastic deformation ability of the compo-
sites during compression.

4. Conclusions

Fig. 7. Compressive stress-strain curves of the Al-Mg2Si composite samples In this study, the microstructure, friction and wear behavior, and
under different solidification conditions. compressive properties of Al–20% Mg2Si composites under different
processing conditions including conventional casting, ST, and high-
Compared with the as-cast and ST composite samples, the sample after pressure-assisted solidification have been compared. Based on the re-
high-pressure-assisted solidification showed extensive plastic deforma- sults, the main conclusions are as follows:
tion. The compression test was stopped at ~ 75% compressive strain
and the samples had no obvious cracks or fractures. The fracture surface (1) The high-pressure-assisted solidification caused not only an in-
of the as-cast samples was relatively flat with the coarse and irregular crease in the eutectic transition temperature of the composite, but
primary Mg2Si phase embedded in the Al matrix, showing limited also an increase in the solubility of Mg and Si in the α-Al matrix,
plastic deformation. Also, cracks were observed in the Mg2Si phase, from 2.85 and 0.61–6.90 and 1.68 wt%, respectively. The high so-
indicating typical cleavage fractures. The ST composite showed a lubility of Mg and Si in the α-Al matrix gave rise to a substantial
fracture surface with more dissociated facets and local fishbone-like solid-solution strengthening of the composite.
patterns (magnified in Fig. 8c), indicating cleavage fractures. The The primary Mg2Si phase that appeared in the conventionally cast
composite after high-pressure solidification showed a river pattern in Al–Mg2Si composite was not observed in the composite after high-
the fractured surface with a few dissociated facets and many torn edges pressure-assisted solidification; in contrast, a primary α-Al phase
without primary phase particles, indicating quasi-cleavage fractures. precipitated preferentially and the grain size of the eutectic Mg2Si
During the compressive deformation of the composite samples, the phase was significantly decreased.
applied load on the composite was transmitted to the particles through (2) The friction coefficient, wear loss, and surface roughness of the
interfaces. The as-cast composite samples consisted of a coarse, irre- composite after high-pressure assisted solidification substantially
gular primary Mg2Si phase and a coarse, fibrous eutectic Mg2Si phase. decreased. The friction coefficient of the composite under high-
During the plastic deformation process, the primary phase was a non- pressure solidification was only 44% of that of the as-cast composite
deformable particle phase; thus dislocation lines could only bypass the and 67% of that of the ST composite.
particles. Therefore, it was difficult for the particles to synergistically (3) The hardness, compressive yield strength, and plastic deformation
deform with the composite matrix during deformation, which then led ability of the composite after high-pressure-assisted solidification
to stress concentration and resulted in cracking and/or peeling of the increased significantly. The fracture mode of the composite after
primary particles from the matrix at brittle fractures. Therefore, the compression testing transformed from cleavage fractures in the as-
composites showed low compression strength and deformation ability. cast composite to quasi-cleavage fractures in the high-pressure-so-
After ST and high-pressure-assisted solidification, the compressive yield lidified composite.
strength and plastic deformation ability of the composites were sig- (4) The modes of wear were converted from adhesive wear and abra-
nificantly enhanced. According to Fleischer's formula [28], the yield sive wear in the as-cast state to only abrasive wear after high-

Fig. 8. SEM images of fracture morphology of Al-Mg2Si composite samples under different solidification conditions: (a) as-cast, (b) ST, (c) high pressure of 3 GPa
assisted.

14
X. Tong et al. Materials Science & Engineering A 733 (2018) 9–15

pressure-assisted solidification. [9] J.T. Zhang, Y.G. Zhao, X.F. Xu, X.B. Liu, Trans. Nonferr. Met. Soc China 23 (2013)
2852–2856.
[10] G.Z. Bai, Z. Liu, J.X. Lin, Z.F. Yu, Y.M. Hu, C. Wen, Mater. Des. 90 (2016) 424–432.
Acknowledgments [11] J.F. Jiang, Y. Wang, Mater. Sci. Eng. A 639 (2015) 350–358.
[12] M. Emamy, S.E. Vaziri Yeganeh, A. Razaghian, K. Tavighi, Mater. Sci. Eng. A 586
This work was supported by the Department of Education of (2013) 190–196.
[13] H.Y. Wang, L. Liu, Y. Chen, J. Zhao, J.H. Liu, R.J. Zhang, Chin. J. High Press. Phys.
Zhejiang Province for financial support (Y201432105). CW and YL also 27 (2013) 768–772.
acknowledge the financial support for this research by the Australian [14] P. Ma, C.M. Zou, H.W. Wang, S. Scudino, B.G. Fu, Z.J. Wei, U. Kühn, J. Eckert, J.
Research Council (ARC) through the Discovery Project (DP170102557) Alloy. Compd. 586 (2014) 639–644.
[15] J.C. Jie, C.M. Zou, H.W. Wang, Z.J. Wei, Mater. Lett. 64 (2010) 869–871.
and the Future Fellowship (FT160100252). JL acknowledges the fi- [16] R. Xu, H. Zhao, J. Li, R. Liu, W.K. Wang, Mater. Lett. 60 (2006) 783–785.
nancial support for this research by the Scientific Research Fund of [17] Y.S. Han, D.H. Kim, H.I. Lee, Y.G. Kim, Scr. Metal. Mater. 31 (1994) 1623–1628.
Hunan Provincial Science and Technology Department through the [18] Y. Dong, X.P. Lin, R. Xu, R.G. Zheng, Z.B. Fan, S.J. Liu, Z. Wang, Mater. Lett. 32
(2014) (1078-1055).
projects 2018JJ4053, 2016GK4035 and 2016JC2005.
[19] H. Wang, D. Zhu, C. Zou, Z. Wei, Mater. Des. 34 (2012) 488–493.
[20] P.W. Fish, Phys. Educ. 6 (1971) 7.
References [21] M.F. Zhu, L. Zhang, HongL. Zhao, Doru M. Stefanescu, Acta Mater. 84 (2015)
413–425.
[22] A. Jayakaman, W. Klement Jr., R.C. Newton, G.C. Kennedy, J. Phys. Chem. Solids
[1] Q.D. Qin, W.X. Li, K.W. Zhao, S.L. Qiu, Y.G. Zhao, Mater. Sci. Eng. A 527 (2010) 24 (1963) 7–18.
2253–2257. [23] J.C. Jie, C.M. Zou, H.W. Wang, B. Li, Z.J. Wei, J. Alloy. Compd. 510 (2012) 11–14.
[2] Q.X. Jing, C.X. Zhang, X.D. Huang, China Foundry 2 (2009) 133–136. [24] J. Zhang, Z. Fan, Y.Q. Wang, B.L. Zhou, J. Mater. Sci. Technol. 17 (2001) 494–496.
[3] M. Emamy, H.R. Jafari Nodooshan, A. Malekan, Mater. Des. 32 (2011) 4559–4566. [25] R.D. Li, X.S. Cao, Y.D. Qu, Y. Xie, R.X. Li, C. Tian, Trans. Nonferr. Met. Soc. China
[4] N.A. Nordin, S. Farahany, A. Ourdjini, T.A.A. Bakar, E. Hamzah, Mater. Charact. 86 19 (2009) 1570–1574.
(2013) 97–107. [26] X.B. Liu, Y.G. Zhao, Acta Metall. Sin. 50 (2014) 753–761.
[5] E. Georgatis, A. Lekatou, A.E. Karantzalis, H. Petropoulos, S. Katsamakis, A. Poulia, [27] J. Zhao, L. Liu, J.R. Yang, G.R. Peng, J.H. Liu, Rj Zhang, G.Z. Xing, Rare Met. 27
J. Mater. Eng. Perform. 22 (2013) 729–741. (2008) 541–544.
[6] H.C. Yu, H.Y. Wang, L. Chen, M. Zha, C. Wang, C. Li, Q.C. Jiang, Mater. Sci. Eng. A [28] R.L. Fleischer, Acta Metall. 11 (1963) 203–209.
685 (2017) 31–38. [29] P. Ma, Z.J. Wei, Y.D. Jia, C.M. Zou, S. Scudino, K.G. Prashanth, Z.S. Yu, S.L. Yang,
[7] S.C. Ram, K. Chattopadhyay, I. Chakrabarty, J. Alloy. Compd. 724 (2017) 84–97. C.G. Li, J. Eckert, J. Alloy. Compd. 688 (2016) 88–93.
[8] X.D. Lin, C.M. Liu, Y.B. Zhai, K. Wang, J. Mater. Sci. 46 (2011) 1058–1075.

15

S-ar putea să vă placă și