You are on page 1of 8

Progress in Natural Science: Materials International 30 (2020) 221–228

HOSTED BY Contents lists available at ScienceDirect

Progress in Natural Science: Materials International


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

Original Research

On the characterization of microstructure and fracture in a high-pressure T


die-casting Al-10 wt%Si alloy
Xiangyi Jiaoa, Chaofeng Liua, Jun Wanga, Zhipeng Guoa,b,∗∗, Junyou Wangc, Zhuoming Wangc,
Junming Gaoc, Shoumei Xionga,b,∗
a
School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
b
Key Laboratory for Advanced Materials Processing Technology, Tsinghua University, Beijing China
c
Hong Bang Die Casting (NanTong) CO.,LTD China

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

Keywords: Microstructure and fracture behavior in a high-pressure die-casting Al-10 wt%Si alloy have been investigated
Fracture using optical microscope (OM), scanning electron microscope (SEM) and a high-resolution laboratory computed
Hypoeutectic Al-10Si tomography (CT). The results showed that a typical heterogeneous microstructure of the alloy comprised α-Al
Microstructure heterogeneity rich region, eutectic silicon band region and porosity. The microstructure patterns highly dependent on fluid
High-pressure die-casting
convection and rapid solidification. Under high filling speed, externally solidified crystals (ESCs) and the
growing dendrites migrated in center and formed α-Al rich region. Si particles was discharged and enriched in
the final solidified liquid, forming eutectic silicon band. Hard Si particles and brittle Fe-rich phases served as
obstacles prevented dislocation migration, causing local stress concentration. Due to large movable slip systems
in α-Al rich region, the propagation path of the crack was greatly extended. Net-shrinkage that induced by dense
impinging dendrites led to the microcracks along the boundary of dendrites which promoted intergranular
fracture.

1. Introduction geometry, size and distribution of the porosity. Niklas et al. [8] con-
sidered that the existence of gas porosity made it difficult to apply heat
Hypoeutectic Al-10 wt%Si alloys are extensively applied in auto- treatment in die casting because surface blistering occurred during
motive industry due to low density, good formability and high specific solution and heat treatment. Most studies have been performed to in-
strength [1,2]. Structural components manufactured using Al-10 wt%Si vestigate the effect of ESCs and defect band on the casting performance
alloys have promising mechanical properties because of the narrow [9]. The macroscopical defects, such as, large shrinkages network in a
solidification freezing range. Zhang et al. [3] found that Al–10- casting, normally served as the crack initiation source. These defects
Si–1.2Cu–0.7Mn exhibited excellent tensile properties, such as, the YS can induce stress concentration followed by crack initiation and pro-
of 206 MPa, the UTS of 331 MPa and the elongation (EI) of 10%. pagation when the external stress exceeded the yield strength. Using in-
High pressure die casting (HPDC) is one of the most popular situ tensile test, Li et al. [10] found that the specimen was fractured by
methods to produce complex thin-wall components [4]. However, the connecting defects at the section with minimum effective force bearing
rapid filling and fast solidification can lead to the formation of parti- area. In addition, the interface between Al and Si also provided the path
cular microstructure. The so-called externally solidified crystals (ESCs) for crack propagation. Lee et al. [11] found that the microporosity was
normally form in the shot sleeve during the slow-shot stage. The defect induced at the phase interface between Al and eutectic Si particles
band normally exhibited in HPDC components and its formation was during deformation. These microporosity would then act as crack in-
highly dependent on melt flow [5]. In addition, massive pores existed in itiation source and promote crack expansion, similar finding in Refs.
castings, which was affected by the air entrapment in shot sleeve. [12–14].
Outmani et al. [6] found that the porosity level was a key factor to This paper investigates the relation of the microstructure and frac-
decrease elongation of Al-Si-Cu HPDC castings. Lu et al. [7] found that ture of a high-pressure die-casting Al-10 wt%Si alloy. The 3-D large size
the initiation and expansion of the fatigue crack were highly sensitive to reconfiguration of fracture morphology was proposed using high-


Corresponding author. School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China.
∗∗
Corresponding author. School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China.
E-mail addresses: zhipeng_guo@mail.tsinghua.edu.cn (Z. Guo), smxiong@tsinghua.edu.cn (S. Xiong).

https://doi.org/10.1016/j.pnsc.2019.04.008
Received 31 January 2019; Received in revised form 25 April 2019; Accepted 28 April 2019
Available online 05 April 2020
1002-0071/ © 2020 Chinese Materials Research Society. Published by Elsevier B.V. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/BY-NC-ND/4.0/).
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

Table 1
Chemical composition of AlSi10MnMg hypoeutectic Al-Si alloy (wt. %).
Si Fe Cu Mn Mg Zn V Ti Sr Al

9.5–11.5 < 0.15 < 0.03 0.5–0.8 0.1–0.5 < 0.07 < 0.07 0.04–0.15 0.01–0.025 Bal.

Tensile tests were performed on a WDW electronic universal testing


machine with a loading rate of 2 mm/min. Standard HPDC test bar (see
Fig. 1b) was prepared to determine stress-strain curve and three dog
bone bars with the same processing parameters and the same position
(see Fig. 1a) were employed to ensure the accuracy of the tensile test
results. The elongation was recorded by a knife-edge extensometer fixed
in the range of gauge length. All tests were performed at ambient
temperature.

3. Results

3.1. Microstructure heterogeneity

Fig. 2 shows the optical micrograph of the rod with different slow
shot speeds for the HPDC AlSi10MnMg alloy. Particular microstructure
patterns including α-Al rich region and eutectic silicon band were
present in Fig. 2a and b, namely, α-Al enriched in central region while
eutectic silicon-rich region surrounded the α-Al rich region with a ring
shape. A large volume fraction of pores (gas pore and shrinkage) were
exhibited across the sample.
In slow shot sleeve, metal liquid turbulence was caused by the
plunger movement enforced more gas in the casting. These gases were
sealed to form large size gas pores. Below a circular cross section, the
Fig. 1. Configuration of the casting and sample positions applied in the HPDC
square cross section shows the cutaway view along height direction.
experiment. (a) the casting including three tensile test bars and one plate; (b)
The finding confirms that α-Al rich region and eutectic silicon-rich re-
the sketching of tensile test bar.
gion were consistent with the results. The spacing of two eutectic si-
licon band was about 3 mm. The thickness of each eutectic silicon band
resolution computed tomography (CT) equipped with a software was about 500 μm. According to Ref. [16], the position and thickness of
namely Avizo. the eutectic segregation bands depended on thermal gradients ( ΔT ) as
well as the solid fraction gradient ( f sESCs ). In other words, the eutectic
2. Experimental procedures segregation bands were changeable affected by die temperature and
ESCs content. In Fig. 2b, the particular α-Al rich region marked by red
A commercial hypoeutectic aluminum alloy, AlSi10MnMg, was frame existed and it even appeared out the eutectic silicon band.
used, and its chemical composition is shown in Table 1. The liquidus Fig. 3 shows the microstructure of α-Al rich region and eutectic
and solidus temperature of this alloy was 595 °C and 550 °C, respec- silicon band region. In α-Al rich region (Fig. 3a and b), the micro-
tively. Fig. 1a shows the configuration of the specific casting including structure comprised ESCs, fine α-Al grains, eutectic. Coarse ESC with
three tensile test bars and one plate. Standard tensile testing bars with a the size over 30 μm was present when VL = 0.1 m/s, while fine ESCs
diameter of 6.4 mm (see Fig. 1b) were produced by a TOYO BD-350V5 with average grain size about 10 μm when VL = 0.4 m/s. In addition,
cold chamber die casting machine, the detail of which can be found in the shrinkage with a size about 2 μm formed along grain boundary and
the previous study [15]. The sample location for OM and SEM analysis developed into network in Fig. 2b. Contrary to the formation of the gas
was marked in Fig. 1b. Table 2 shows the key processing parameters pores, the shrinkage was caused by feeding failure in impinging den-
adopted in HPDC experiment. Two samples with different slow shot drites.
speeds (VL) were produced to obtain different HPDC microstructure. Fig. 3c and d shows the eutectic silicon-rich region of HPDC Al-
In metallography observation of the microstructure, SiC papers of Si10MnMg at different slow shot speeds. The eutectic silicon band zone
#400, #800, #1200, #1500 and #2000 were used in order. After comprised Fe-rich phase and eutectic silicon phase. Fe-rich phase be-
mechanically mounted, ground, polished and ultrasound cleaning, the haved in thick needle-plate shape, while eutectic silicon was more fiber-
microstructure was observed using OM and SEM. A software, namely and rod-like. Since the preferred growth orientation of Si particles
Image-Pro Plus 6.0, was used to identify phases in samples. A high- was < 110 > or < 211 > located on {111} planes, Si particles ex-
resolution laboratory CT, Nanotom m manufactured by GE, was used to hibited a facet characteristic with sharp edges. 0.02 wt% Sr in this alloy
study the 3-D fracture morphology in samples. The voltage and electric was gathered in twins and dislocations of Si particles, inhibiting the
current were set to be 110 kV and 100 μA, respectively. growth of Si particles. Fine silicon particles were preserved, as shown in

Table 2
Processing parameters during HPDC.
Sample Melting temperature (°C) Initial mold temperature (°C) Intensification casting pressure (MPa) Slow shot speed (m/s) Fast shot speed (m/s)

1 680 120 87 0.1 2.75


2 680 120 87 0.4 2.75

222
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

Fig. 2. Optical micrograph of cross-section rod in (a) VL = 0.1 m/s, (b) VL = 0.4 m/s.

Fig. 3c and d. Fig. 4a. When the melt was injected into the die cavity, the precipitation
Fig. 4 shows the type of Fe-rich phases in HPDC AlSi10MnMg alloy. of Fe affected by high cooling rate was restricted, and the final solidi-
Two types of Fe-rich phases, including bulk Fe-rich phase and lath Fe- fication of Fe was at the eutectic interface, to form δ-Al4FeSi2 (lath Fe-
rich phase, could be identified. The bulk Fe-rich phase distributed rich phase), as shown in Fig. 4b. Because of massive precipitation of β-
random, and its size varied considerably. The lath Fe-rich phase was Al5FeSi in HPDC initial stage, higher Mn content (25.59%) and Fe
surrounded by eutectic and segregated in strip shape. According to Al- content (3.62%) were in bulk Fe-rich phase. Nevertheless, higher Si
Si-Fe diagram [17], β-Al5FeSi nucleated preferentially in liquid phase content (∼13.9%) was in lath Fe-rich phase because of the final soli-
and further expanded to form large flake-shape under low cooling speed dification combined with eutectic.
of ∼1 °C/min. In die-casting process, β-Al5FeSi preferentially formed in
slow shot sleeve under lower cooling rate, and then the flake β-Al5FeSi
was broken affected by high speed fluid, to form bulk Fe-rich phase in

Fig. 3. (a), (b) α-Al rich region, and (c), (d) eutectic silicon-rich region, of HPDC AlSi10MnMg at (a) and (c) VL = 0.1 m/s, (b) and (d) VL = 0.4 m/s.

223
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

Fig. 4. (a) Bulk Fe-rich phase, (b) lath Fe-rich phase, in HPDC AlSi10MnMg alloy.

Table 3 4. Discussion
Mechanical properties of HPDC AlSi10MnMg alloy.
Sample YS(MPa) UTS(MPa) EI(%)
4.1. The formation of heterogeneous microstructure

1 157 ± 2 287.4 ± 14.6 5.8 ± 0.14 In the process of die-casting, the solidification contained two stages.
2 168 ± 3.6 307.9 ± 6.9 4.6 ± 0.62 In the first stage, ESCs nucleated in slow shot sleeve and grew with the
motion of the plunger. Lower slow shot speed would extend the dura-
tion of the melt in the shot sleeve, and increase the size and area
3.2. Fracture feature
fraction of ESCs. In Fig. 7a, coarser ESCs with the shape of dendrite was
present due to the longer stay time in shot sleeve. At the end of the first
Table 3 lists the mechanical properties of HPDC AlSi10MnMg alloy.
stage, ESCs and liquid formed semi-solid state. The second stage hap-
In comparison with the casting under lower slow shot speed (0.1 m/s),
pened in die cavity. The cooling rate in die cavity was much higher,
Sample 2 with higher slow shot speed of 0.4 m/s exhibited the im-
such as the interfacial heat transfer coefficient between the metal and
proved strength, while the decreased elongation, such as, the higher YS
die in die cavity could approach ∼104–105 W m−2 K−1 and the cooling
of∼7%, the higher UTS of ∼7%, the lower EI of ∼21%.
rate could be ∼100 K s−1. Therefore, the nucleation rate was much
Fig. 5a and b shows the whole crack surface with high field depth.
high in metal liquid, which restricted the growth of α-Al grains. When
Compared with that in Fig. 2, α-Al rich region and eutectic silicon band
the metal liquid was injected in die cavity, explosive nucleation of α-Al
were marked in Fig. 5. Caused by the connected shrinkages, penetrated
grains happened in remaining liquid, forming large quantity fine α-Al
porous inclusions existed in center, to cause the rough fracture surface
grains, as shown in Fig. 7. In addition to rapid solidification, the high
and tortuosity of the crack path. In eutectic band zone, the fracture
speed liquid flow also contributed to the heterogeneous microstructure.
surface was relatively flat. Fig. 6a and b shows the whole 3-D fracture
Due to the disperse effect of high filling speed, coarse ESCs distributed
morphology contour map of the undulation height of the fracture sur-
in every position along the radius direction. However, relative dense
face with 3700 slices piled up. The half of total crack fluctuation is
distribution of ESCs was in center to form α-Al rich region. According to
defined as the baseline to determine the fracture level. Based on the
the previous study [18], the ESCs and growing dendrites tended to
level of crack fluctuation, the 3-D fracture surface was classified into
migrate toward the center instead of attaching to the solidification layer
ridge (red), basin (blue) and plain (green) areas. Large crack fluctuation
during filling. The eutectic segregation band was presumed to come
existed at edge. The cracks were relatively flat at the center of sample.
from the last solidified liquid. According to Li et al. [19], this segre-
Since deep central pit and tortuous crack surface exhibited in center
gation band followed the contour of melt flow in liquid filling, which
when VL = 0.4 m/s, it indicates a longer crack propagation path. The
can contribute to the feeding in interdendritic gap.
rise and fall of the flat propagation surface was measured to be 1320 μm
The dendrites growing along the wall were fused and dissociated in
when VL = 0.1 m/s. Because of large volume of shrinkage in center, the
center. Equipped with ESCs, α-Al rich region formed in center. While
whole height of crack fluctuation was highly improved and reached
the compact chilled crystal formed in surface affected by chill effect of
2064 μm when VL = 0.4 m/s. When the main crack passed through the
die cavity [20–22]. The middle region between α-Al rich region and
shrinkages region, the crack propagation direction changed along
surface chilling shell provided the path for metal liquid filling. Because
shrinkage network. Accordingly, the crack fluctuation increased. The
of massive precipitation of α-Al, the last solidified liquid in middle
microcracks around the shrinkage converged on the main failure crack,
region would have high silicon concentration. Therefore, the eutectic
and propagated together with the main crack.

Fig. 5. Whole crack surface of (a) VL = 0.1 m/s and (b) VL = 0.4 m/s.

224
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

Fig. 6. 3-D fracture morphology and contour map of the undulation height of the fracture surface for (a) VL = 0.1 m/s and (b) VL = 0.4 m/s.

silicon band formed. The pores in HPDC aluminium alloy comprised gas pores and
shrinkage pores. Gas pores were induced by the air entrapment in shot
sleeve, while shrinkage pore were induced by the feeding failure during
4.2. Effect of pores and Fe-rich phases on the fracture
solidification. In Fig. 8a, the shape of gas pore was nearly spherical and
its size was about 60 μm. However, the size of shrinkage pore was
The main Cracks originated from the surface and changed direction
smaller and less than 5 μm. The shrinkage pore distributed along the Al
in the course of expansion. The second cracks were constantly sprouted
grain boundary and its shape was narrow ribbon due to the solidify
up during the expansion process.

Fig. 7. Distribution of ESCs along the radius direction at (a) VL = 0.1 m/s; (b) VL = 0.4 m/s.

225
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

Fig. 8. Typical defects and fracture (a), (d) gas pore, (b), (e) shrinkage pore, and (c), (f) Fe-rich phase.

Fig. 9. (a), (c) Eutectic silicon rich region, (b), (d) α-Al rich region, at (a), (b)VL = 0.1 m/s, (c), (d) VL = 0.4 m/s, and (e) quantity statistics of α-Al grain size.

contraction. The worse roundness of shrinkage was easy to cause stress of sample and contributed to the failure of tensile test bar. Because of
concentration, leading to microcrack initiation around its tip. Ac- natural brittle property and poor deformation compatibility, the bulk
cording to Ref. [23], the increase of porosity increased the degree of Fe-rich phase was broken into several pieces, as shown in Fig. 8f. Ac-
maximum stress (σmax ) . The expression was as follow: cording to Zhu et al. [25], the coarse Fe-rich phases inhibited the dis-
location motion and acted as stress raisers to weaken the coherence.
KF
σmax = Meanwhile, the local stress concentration made brittle Fe-rich phases
A0 (1 − f ) (1) easy to crack along the definite crystallographic structural planes.
where K was the stress concentration factor of the porosity, which was
related to the size and shape of pores; F was tensile force; A0 was cross 4.3. Effect of eutectic silicon-rich region and α-Al rich region on fracture
section area; f was the volume fraction of pores. The gas pore with large expansion
volume fraction increase the f while the shrinkage with irregular shape
increase the K. Compared with gas pore, the effect of shrinkage with The hindering effect of the eutectic silicon on dislocation motion
smaller size on fracture could be ignored, but the shrinkage was in- determines the strength of hypoeutectic Al-Si alloy. According to
terconnected in α-Al rich region forming large shrinkage network. It Orowan mechanism, the smaller the particle spacing (d ) and the larger
was proposed by Li et al. [24] and named net-shrinkage. By 3-D CT the volume fraction ( f ), the more obvious the strengthening effect is.
image reconstruction, they concluded that net-shrinkage was the The expressions were as follow:
dominant sources for crack initiation because of a much lower spheri- Gb
city causing high stress intensity. Microcracks tended to originated Δτ ≈
d (2)
around worse roundness of the shrinkage and further expanded along
the boundary of the dendrites. Seen from Fig. 5a and b, net-shrinkage 1
Δτ ≈ αf 2 r −1 (3)
region was located in center and covered the whole α-Al rich region.
The fracture of net-shrinkage became rough, as shown in Fig. 8e, in- where b was burgers vector; r was particle size; α depended on the type
dicating the microcracks covering Al dendrite boundary and a long of dislocation. As shown in Fig. 9a and c, Sample 2 with higher density
crack propagation path. of eutectic silicon owns smaller particle spacing and larger volume
In Fig. 8c, the bulk Fe-rich phase distributed random in cross section fraction. Table 3 provides the reason for higher strength of Sample 2.

226
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

However, when the dislocations encounter with the obstacles e.g. the micropores existed in the sample, as shown in Fig. 10a. As for lath Fe-
eutectic silicon or Fe-rich phase, the motion of the dislocations will be rich phase, it was easy to be broken affected by the nature brittle, and
stopped. The dislocations piled up along {111} planes can lead to stress could not provide the effective strength for inhibiting crack propagation
concentration in phase interface, and then microcracks begin to ger- [29]. High stress intensity around these defects would occur, which
minate and extend along the interface of Al and Si. According to Lee promoted the generation and propagation of the microcracks.
[26], the eutectic silicon rotated to form micropores because of in- The α-Al rich region had also an important influence on the ex-
congruity of deformation between Al and Si in tensile test. The weak pansion of the crack. Coarse ESCs were soft and easy to deform. These
incoherent interface provided a path for the crack growth. grains acted as a stress buffer, if the cracks passed the interior of the
Many researchers have found that part of eutectic silicon is ruptured grain. When the energy was consumed and the crack growth was vo-
in tensile fracture. The broken particles generate new surfaces and in- latile affected by alternating sliding {111} < 110 > , according to the
crease the interfacial energy. The increased critical shear stress ( Δτa ) FCC crystal structure. The continuous change of the lattice plane and
could be expressed: crystal direction during sliding was the key factor to increase the crack
3 1 propagation path. As shown in Fig. 9b and d, the coarse ESCs and fine
1.1 σ 2 f 2 1 α-Al grains were common existence in center region. Higher density of
Δτa = r2
α Gb2 (4) ESCs in Sample 1 would increase the deformability. However, the
denser and smaller α-Al in Sample 2 increased the density of boundaries
where σ was interface energy. Similarly, the smaller size (r) and large
and provided more sites for shrinkages. Fig. 9e shows the quantity
fraction of eutectic silicon will increase extra strength. Man [27] found
statistics of α-Al grains size from 0 μm to 120 μm. The grain sizes less
that the fracture Si particles increased with stress improvement, and Si
than 8 μm were account for the majority. The maximum grain size in
particles with large length-width ratio were easier to fracture. Wang
Sample 1 was 108 μm, far greater than that in Sample 2 (92 μm). The
[28] discovered the probability ( p ) of silicon fracture to be conformed
largest size of ESCs was typical dendritic crystal, as shown in Fig. 9e. In
to Weibull:
the crack surface in Fig. 10c and d, the dendrites and tearing dendrites
m
V σ presented, which was related to intergranular fracture and transgra-
p= 1 − exp ⎡− ⎛ P ⎞ ⎤ ⎜ ⎟
⎢ Vo ⎝ σo ⎠ ⎥ (5) nular fracture, respectively. Transgranular fracture surface with small
⎣ ⎦
dimple and tearing ridge was rough while intergranular fracture surface
where m was Weibull modulus; V and σP were particle volume and the was smooth. Net-shrinkage induced by the impinging dendrites was the
actual force applied in particle, respectively. Vo and σo were related to reason for the intergranular crack propagation. Because of dense α-Al
microstructure of hypoeutectic Al-Si alloy. The greater of V and σP dendrites in Sample 2, a large amount of shrinkages increased the
value, the fraction silicon fracture will be higher. Since denser Si par- possibility of intergranular fracture. Meanwhile, the edge of shrinkage
ticles distribution in Sample 2 lead to more broken Si particles and endured the high local stress intensity, which was easy to induce mi-
higher Δτa , the overall strength is therefore improved. crocracks, and promotes the crack growth.
In Fig. 10a and b, the quasi-cleavage fracture morphology consists
of tearing edges, micropores and the exposed lath Fe-rich phases, which 5. Conclusions
was present in eutectic silicon band region. Meanwhile, the large
cleavage steps formed, indicating the change of main crack propagation (1) Particular heterogeneous microstructure, including α-Al rich region
direction. In the process of main crack propagation, when the tip of the and eutectic silicon band region, exhibit in the HPDC AlSi10MnMg
crack was in contact with the eutectic silicon band region, the growth of alloy.
cracks was hindered by dense eutectic silicon particles, which improved (2) Bulk Fe-rich phases and lath Fe-rich phases exist in
the strength of alloy. However, many broken Fe-rich phase and

Fig. 10. Fracture morphology of (a), (b) eutectic silicon rich region, (c), (d) α-Al rich region at (a), (c) VL = 0.1 m/s, and (b), (d) VL = 0.4 m/s.

227
X. Jiao, et al. Progress in Natural Science: Materials International 30 (2020) 221–228

HPDCAlSi10MnMg alloy. All brittle phases can contribute to crack [5] X. Li, Z. Guo, S. Xiong, Mater. Char. 129 (2017) 344–352.
propagation. [6] I. Outmani, L. Fouilland-Paille, J. Isselin, M. El Mansori, J. Mater. Process. Technol.
249 (2017) 559–569.
(3) Dense silicon particles in eutectic silicon band region inhibit the [7] Y. Lu, F. Taheri, M.A. Gharghouri, H.P. Han, J. Alloys Compd. 470 (2009) 202–213.
crack propagation, while coarse ESCs and dense α-Al in center in- [8] A. Niklas, S. Orden, A. Bakedano, M. Da Silva, E. Nogués, A.I. Fernández-Calvo,
crease the crack propagation path. Mater. Sci. Eng. A 667 (2016) 376–382.
[9] X. Li, S.M. Xiong, Z. Guo, Mater. Sci. Eng. A 674 (2016) 687–695.
(4) Interdendritic shrinkages connect to each other to form the net- [10] X. Li, S.M. Xiong, Z. Guo, Mater. Sci. Eng. A 672 (2016) 216–225.
shrinkage in α-Al rich region. This net-shrinkage is the dominant [11] C. Lee, J. Youn, Y. Lee, Y. Kim, Mater. Sci. Eng. A 678 (2016) 227–234.
sources for microcrack initiation and propagation. [12] J. Wang, Z. Guo, J.L. Song, W.X. Hu, J.C. Li, S.M. Xiong, Mater. Des. 137 (2018)
176–183.
[13] J. Wang, Z. Guo, S.M. Xiong, Mater. Char. 123 (2017) 354–359.
Acknowledgement [14] A.D. Klarner, J. Miao, W. Sun, A.A. Luo, X. Zeng, Metall. Mater. Trans. A 50 (2019)
1522–1533.
[15] X. Li, S.M. Xiong, Z. Guo, J. Mater. Process. Technol. 231 (2016) 1–7.
The authors would like to thank the National Natural Science
[16] C.M. Gourlay, A.K. Dahle, H.I. Laukli, Metall. Mater. Trans. A 35 (2004)
Foundation of China (Grant No. 51775297), the National Science and 2881–2891.
the Tsinghua University Initiative Scientific Research Program [17] S. Ji, W. Yang, F. Gao, D. Watson, Z. Fan, Mater. Sci. Eng. A 564 (2013) 130–139.
(20151080370) and UK Royal Academy of Engineering/Royal Society [18] C.M. Gourlay, A.K. Dahle, H.I. Laukli, Metall. Mater. Trans. A 35 (2004)
2881–2891.
through the Newton International Fellowship Scheme. [19] X. Li, S.M. Xiong, Z. Guo, Mater. Sci. Eng. A 633 (2015) 35–41.
[20] M. Yang, S.M. Xiong, Z. Guo, Acta Mater. 112 (2016) 261–272.
Appendix A. Supplementary data [21] Z. Guo, S. Xiong, S.H. Cho, J.K. Choi, Acta Metall. Sin. (2007) 1149–1154.
[22] J. Wang, Z. Guo, X.Y. Jiao, S.M. Xiong, Mater. Char. 140 (2018) 179–188.
[23] L. Zuo, B. Ye, J. Feng, X. Kong, H. Jiang, W. Ding, J. Mater. Sci. Technol. 34 (2018)
Supplementary data to this article can be found online at https:// 1222–1228.
doi.org/10.1016/j.pnsc.2019.04.008. [24] X. Li, S.M. Xiong, Z. Guo, J. Mater. Sci. Technol. 32 (2016) 54–61.
[25] X. Zhu, P. Blake, K. Dou, S. Ji, Mater. Sci. Eng. A 732 (2018) 240–250.
[26] C. Lee, J. Youn, Y. Lee, Y. Kim, Mater. Sci. Eng. A 678 (2016) 227–234.
References [27] M. Zhu, Z. Jian, L. Yao, C. Liu, G. Yang, Y. Zhou, J. Mater. Sci. 46 (2011)
2685–2694.
[28] Q.G. Wang, C.H. Caceres, J.R. Griffiths, Metall. Mater. Trans. A 34 (2003)
[1] K. Wang, H.Y. Jiang, Q.D. Wang, B. Ye, W.J. Ding, Mater. Char. 117 (2016) 41–46.
2901–2912.
[2] S. Nafisi, R. Ghomashchi, H. Vali, Mater. Char. 59 (2008) 1466–1473.
[29] Z. Li, N. Limodin, A. Tandjaoui, P. Quaegebeur, P. Osmond, D. Balloy, Mater. Sci.
[3] P. Zhang, Z. Li, B. Liu, W. Ding, J. Mater. Sci. Technol. 33 (2017) 367–378.
Eng. A 689 (2017) 286–297.
[4] X.J. Wang, L.C. Zhang, M.H. Fang, T.B. Sercombe, Mater. Sci. Eng. A 597 (2014)
370–375.

228

You might also like