J Msea 2012 02 073

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Materials Science and Engineering A 543 (2012) 185–192

Contents lists available at SciVerse ScienceDirect

Materials Science and Engineering A


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

A study of the mechanical properties of an Al–Si–Cu alloy (ADC12) produced by


various casting processes
M. Okayasu a,∗ , Y. Ohkura b , S. Takeuchi a , S. Takasu b , H. Ohfuji c , T. Shiraishi a
a
Department of Materials Science and Engineering, Ehime University, 3 Bunkyo-cho, Matsuyama, Ehime 079-8577, Japan
b
Department of Machine Intelligence and Systems Engineering, Akita Prefectural University, 84-4 Ebinokuchi, Tuchiya-aza, Yurihonjo-city, Akita 015-0055, Japan
c
Geodynamics Research Center, Ehime University, 2-5 Bunkyo-cho, Matsuyama, Ehime 079-8577, Japan

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

Article history: The mechanical properties of an Al–Si–Cu alloy (ADC12), produced using various casting technologies,
Received 19 January 2012 have been examined experimentally. Four different casting processes were employed, including grav-
Received in revised form 20 February 2012 ity casting (GC), cold-chamber die-casting (CD), twin rolled continuous casting (TRC) and the Ohno
Accepted 21 February 2012
continuous casting process (OCC). Although these produced the same Al–Si–Cu aluminum alloy, dif-
Available online 3 March 2012
ferent mechanical properties were obtained, in particular microstructural characteristics and dislocation
density. The microstructure of GC and CD samples was formed mainly with coarse ˛-Al phase and needle-
Keywords:
shaped Si and Fe based eutectic structures. In contrast, a fine round ˛-Al phase and tiny eutectic structures
Casting
Twin rolled casting
were observed for the TRC and OCC samples. Such a change of microstructure was caused by the differ-
Continuous casting ent casting process parameters, namely injection speed, casting pressure and cooling rate. High internal
Mechanical property stress as well as high dislocation density was detected for GC and TRC, caused by the high shrinkage
Microstructural characteristic force and high applied rolling force, respectively. Because of the different material properties, the tensile
Aluminum alloy and fatigue strength were altered. A clear Hall–Petch relation with  0.2 = ky d−0.5 + B was obtained, and the
fatigue properties were evaluated with the power law dependence  a =  f Nf −b . The mechanical proper-
ties obtained were also analyzed in relation to the crystal orientation and lattice mis-orientation angle.

© 2012 Elsevier B.V. All rights reserved.

1. Introduction blow holes, abnormal microstructures and shrinkage porosities


[1–3]. Note that abnormal microstructures are obtained specifically
Recently, the production of cast aluminum alloys has been in the cold-chamber die-casting samples, such structures being
increasing, and in Japan the total quantity was approximately formed at the interface between coarse and normal microstruc-
1500 kt in 2010. This large tonnage of cast Al alloys has been used for tures. The coarse microstructures are created by solidification in
automotive parts, such as transmission cases, converter housings the shot sleeve. Moreover, due to the high injection speed for
and cylinder blocks. These automotive parts, previously made of die-casting process, blow holes are present in these cast compo-
cast iron, have been replaced with cast aluminum alloys, because nents. In the gravity cast process, a large amount of Al alloy is
of their contribution to higher fuel efficiency and lower levels of used to make a thick cast component, which can sometimes lead
emissions. The Al–Si–Cu alloy ADC12 is a cast aluminum alloy to shrinkage porosity. In a conventional continuous casting pro-
widely used due to its excellent material properties, namely high cess, molten metal is solidified into a cooled hollow mould just
cast-ability, low density, high productivity, low shrinkage rate and before subsequent rolling in the finishing mills. Under these cir-
relatively high strength. However, cast aluminum alloys have not cumstances, equiaxial dendrite growth occurs on the mould wall,
been always suitable for automotive parts. This is because vari- resulting in shrinkage porosity in the middle of the cast compo-
ous cast defects are included in the cast components. In fact, the nent.
application of cast aluminum alloys to the manufacture of criti- The tensile strength and elongation to failure of such cast Al
cal safety parts in automobiles has been considerably restricted. alloys decreases almost linearly with increasing defect size. The
In previous reports, it was shown that there are various cast reduction rate of the tensile properties is also attributed to the
defects in conventional gravity and die-casting samples, e.g., cracks, type of cast defect. The high reduction rate of the tensile proper-
ties was obtained as the abnormal structure and shrinkage porosity
are included in the cast component. To reduce the numbers of cast
∗ Corresponding author. Tel.: +81 89 927 9811; fax: +81 89 927 9811. defects, an attempt has been made to use alternative casting tech-
E-mail address: okayasu.mitsuhiro.mj@ehime-u.ac.jp (M. Okayasu). nologies, e.g., squeezed casting [4], melt drag twin-roll casting [5],

0921-5093/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2012.02.073
186 M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192

Fig. 1. Microstructures of GC, CD, TRC and OCC samples: (a) optical micrographs; (b) EDX analysis taken by scanning for Al, Si, Cu and Fe.

hot-chamber casting [6], semi-solid rheo-casting [7], thixoforming been used to make copper wire for electrical devices, bismuth-tin
[8] and continuous casting technology [9]. wire for thermal fuses and lead free solder [11]. Some researchers
There exists a unique continuous casting technique proposed have investigated the mechanical properties of OCC-Al alloys [12].
by Ohno, known as the Ohno continuous casting (OCC) process. One of the authors has also examined the mechanical properties
The OCC process is a unidirectional solidification method, and this of OCC-Al alloys (AC8A and Al–33%Cu) [13,14]. It appears that the
casting process provides phase control and texture control [10]. OCC-Al alloys have excellent tensile and fatigue properties because
The OCC process is different from conventional continuous cast- of their unidirectional fine microstructure, low concentration of
ing in that molten metal is poured into a heated hollow mould, defects and uniformly oriented lattice structure. Due to the high
rather than one that is cooled. Until now, the OCC procedure has material quality, the OCC technique may be useful for making
M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192 187

automotive parts, although there are no clear reports regarding


ADC12.
The twin-rolled continuous casting (TRC) process is a process
for making high quality casting products [15], in which the cast
sample is solidified with a high cooling rate using chilled rollers.
It is generally considered that the twin-rolled continuous casting
process has the advantages of low cost and high material quality.
As the processing steps can be omitted in the initial breakdown of
a large size ingot, the running costs can be reduced particularly the
investment in energy and the cost of labor. The microstructure and
mechanical properties of magnesium alloy produced by twin-rolled
continuous casting were examined by Chen et al. They reported
that the microstructure of the TRC strip consisted of a dendritic
structure, whilst the eutectics and intermetallic compounds were
located in the interdendritic region. In addition, they reported that
the TRC sample showed higher strength and lower elongation after
sequential warm rolling, while post annealing after warm rolling
induced a decrease of the strength and an increase of elongation
[16]. Another approach was conducted to make a comparison of
the mechanical properties of TRC-Al alloy products with hot-rolled
strips of aluminum alloy, showing that different tensile strength
and yield stress are caused by different strain hardening effects
[17]. For example, the continuously cast sample has higher tensile
Fig. 2. Relationship between SDAS and cooling rate for cast aluminum alloys
strength, but lower yield stress and lower elastic modulus than the [21–25].
hot-rolled strip ones.
Although extensive experiments have been carried out in order
to understand the material properties of the cast aluminum alloys, The main twin-rolled continuous casting conditions were (i)
there seems to be very little consistency in the available data for temperature of the molten Al alloy in the furnace 973 K ± 5 K, (ii)
ADC12. Again, information describing the mechanical properties roller supporting force 30 t, (iii) roller temperature 298 K and (iv)
of ADC12 is especially important because it is a widely employed casting velocity about 0.4 rpm. The TRC sample was designed to
material for many engineering applications. In particular, informa- have dimensions 1300 mm × 130 mm × 7 mm.
tion on the tensile and fatigue properties is an indispensable part A horizontal OCC system was used, based on the conven-
of the design process for its applications. The aim of this study tional OCC device and consisting of a melting furnace, a cylindrical
is, therefore, to investigate experimentally the mechanical proper- graphite displacer block for molten metal level control, a graphite
ties of cast aluminum alloys (ADC12), produced by several casting mould, a graphite crucible, a cooling device and pinch rollers for
technologies, including the TRC and OCC techniques. withdrawal of the cast sample. The molten sample of Al alloy
of about 0.2 kg was introduced into the graphite crucible, and
then fed into the graphite mould continuously through a runner.
2. Experimental procedures The cast sample was 5 mm in diameter. The mould was heated
by the ceramic heaters, and the target mould temperature was
2.1. Materials 898 K ± 3 K, which is a temperature just above (10 K) the liquidus
of the Al–Si–Cu alloy (ADC12). The cast sample was solidified with
The materials selected in the present work are the cast alu- water cooling. Details of the OCC process can be found in Ref. [18].
minum alloy ADC12. The chemical composition of ADC12 is (in The internal defects (shrinkage porosity and blow holes) in all
mass %) Al–Si10.6 –Cu2.5 –Mg0.3 . In this study, test samples were cast samples were investigated by X-ray diffraction, and cast sam-
produced by four different casting processes: (i) gravity casting ples with fewest defects were used in the present examination.
(GC), (ii) cold-chamber die-casting (CD), (iii) twin-rolled contin-
uous casting (TRC) and (iv) Ohno continuous casting (OCC). 2.2. Tensile and fatigue tests
The temperature of the molten aluminum alloy for the gravity
and die-casting processes was set, in both cases, at 973 K ± 5 K. The Tensile and fatigue tests were carried out using an electro-
gravity casting sample was made by using a standard mould for cre- servo-hydraulic system with 10 kN capacity. Round test specimens
ating the metal ingot, namely a thick and wide rectangular block were used with dimensions 2 mm diameter and 4 mm length. The
(600 mm × 90 mm × 40 mm). 5.8 kg of molten alloy was introduced strain value was measured by a commercial strain gauge. The
to the mould. The die-cast sample was made using a cold-chamber tensile test was conducted with a loading speed of 1 mm/min to
882N die-casting machine. The shape of the die-cast sample was a the failure. The fatigue strength of the cast samples was exam-
simple rectangular plate with dimensions 175 mm × 8 mm × 5 mm. ined using the S–N approach that is the relationship between
The mould temperature in the cavity near the gate was approxi- the applied stress and the cycle number to final failure. The
mately 473 K. The injection speed at the gate was approximately cyclic loading with load control was performed with a sinusoidal
40 m/s. The injection shot was conducted immediately after the waveform at a frequency of 30 Hz and stress ratio of 0.05 up to
molten metal was poured into the shot sleeve. 107 cycles.
For twin-rolled continuous casting (TRC), the samples were
produced using a TRC machine fitted with a high frequency induc- 2.3. Microstructural analysis
tion furnace (Furukawa-Sky Aluminum Corp). The twin-rolled
continuous casting process is described clearly in Ref. [15]. The Microstructural observations of the cast samples were carried
molten Al alloy was installed in a vacuum chamber fitted with out using an optical microscope and a scanning electron micro-
a pair of copper rollers of diameter 400 mm and width 300 mm. scope. In addition, in order to understand the microstructural
188 M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192

Fig. 3. Martens hardness and indentation load–depth curves for GC, CD, TRC and OCC samples obtained at the center of the ˛-Al phase.

characteristics in detail, EBSD scans were carried out using a high smaller step size was used. The samples for EBSD were sectioned
resolution electron JSM-7000F microscope (JEOL Ltd.). The main to less than 5 mm thick perpendicular to the casting direction, and
conditions for this analysis were as follows: the sample was tilted the sample surfaces for the observation were polished to mirror
to an angle of 70◦ and an accelerating voltage of 15 kV, beam cur- flatness in a vibropolisher using colloidal silica for about 2 h. Based
rent 5 nA and step size 1 ␮m or 4 ␮m was used. The step size upon the analysis of the EBSD patterns, the crystal orientations of
was determined on the basis of the required magnification for the matrix grains in the cast Al alloys were analyzed using HKL
the microstructural observations, so for a higher magnification a Channel 5 software.
M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192 189

Fig. 4. Tensile properties of GC, CD, TRC and OCC samples.

3. Results and discussion


Fig. 5. Relationship between the 0.2% proof stress and SDAS (d) for the GC, CD, TRC
and OCC samples.
3.1. Microstructural characteristics

Fig. 1 shows micrographs of the cast aluminum alloys produced curves for the cast samples. The load vs. depth was obtained dur-
by the four different casting processes (GC, CD, TRC and OCC). In ing the indentation loading for the hardness measurement. In this
addition, an EDX analysis was carried out for Al, Si, Cu and Fe in all approach, the hardness test was conducted only in the ˛-Al phases.
samples. From Fig. 1, it is clear that the microstructural character- The load–depth relationship between the elastic constant () and
istics are different. The microstructure for the GC and CD samples depth (d), is different depending on the sample. For instance, there
consists of several phases, including the ˛-Al phase and eutectic is a higher  and lower d value for the OCC sample and lower  and
structures. From the EDX analysis, different eutectic structures are the higher d for the GC and CD samples. In addition, different hard-
found including Cu-, Si- and Fe-based ones. Elongated eutectic sil- ness values are obtained. The hardness of the TRC sample (about
icon and Fe-rich impurity crystals are observed in both GC and CD 1150 N/mm2 ) is more than 15% higher that that of the GC and CD
samples although the size of the eutectic structures is different; samples and about 12% lower than that of the OCC one. Such differ-
there are more eutectic structures for the GC sample. On the other ent hardness values can be explained using the grain size (SDAS);
hand, the eutectic structures are tiny and distributed in between the Hall–Petch relationship predicts that the higher the hardness,
the ˛-Al phases in the TRC and OCC samples. The hardness of the the smaller the SDAS. However, there remains the question that
eutectic Si and Fe structures, examined using an ultra-micro Vickers the hardness for the GC and CD samples is almost the same even
hardness tester, are 4109 N/mm2 and 6361 N/mm2 , respectively, although the SDAS for GC (33.3 ␮m) is about three times greater
which are much higher values compared to that for the ˛-Al phase
and the CuAl2 structures (about 800 N/mm2 ) [19]. In this case, due
to the high hardness, the Si and Fe-based eutectic structures could
be brittle [20].
It is also clear from Fig. 1 that the ˛-Al phase characteristics
are different, depending on the casting method. For instance, a tiny
spherical ˛-Al phase (SDAS 5.5 ␮m and 6.0 ␮m) is obtained for the
OCC and TRC samples, but slightly larger (11.3 ␮m) for the CD sam-
ple. On the other hand, large ˛-Al grains (SDAS = 33.3 ␮m) are seen
in the GC samples. The different SDAS is caused by the different
cooling rates, so that the high cooling rate for TRC and OCC gives
rise to tiny ˛-Al grains. Fig. 2 indicates the relationship between
the SDAS and cooling rates obtained from previous experimental
data [21–25]. From Fig. 2, the cooling rates (CR) for TRC and OCC can
be estimated using the relationship CR = 2 × 104 SDAS−2.67 to give
255 K/s for TRC and 327 K/s for OCC. It should be noted that the ˛-Al
phases for the TRC sample seem to be severely strained. This might
be caused by the 30 t rolling process, as mentioned previously. The
microstructural morphology for the TRC and OCC samples is rela-
tively similar to that for the hot-chamber die-casting sample [6].
The mechanical properties of the hot-chamber die-cast samples
appear to show high tensile strength and high fatigue strength.
As there are different microstructural characteristics in our
cast samples, there would be expected to be different mechanical
properties. Fig. 3 shows the micro-hardness data and load–depth Fig. 6. S–N curves for the GC, CD, TRC and OCC samples.
190 M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192

Fig. 7. (a) Crystal orientation map with pole figures for 0 0 1, 1 0 0 and 1 1 0 poles; (b) misorientation profile in a grain for the GC and CD samples.
M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192 191

Fig. 8. SEM images of fracture surfaces obtained by fatigue testing of GC and OCC samples.

than for CD (11.3 ␮m) (see Fig. 2). The reason behind this will be failure,  a =  f Nf −b , where  a is the stress amplitude in MPa,  f is
discussed in a later section of this paper. the fatigue strength coefficient, Nf is the cycle number to failure
and b is the fatigue exponent. In this case, a high fatigue strength
is expected for a high fatigue strength coefficient  f . The values of
3.2. Mechanical properties  f and b for the cast samples, obtained by least squares analysis,
are  f = 594 MPa and b = −0.07 for the TRC sample and  f = 630 MPa
Fig. 4 shows the tensile properties of the cast samples. It is clear and b = −0.07 for the OCC sample. The  f value for the OCC sample
that the ultimate tensile stress  UTS and strain to failure εf for the is relatively close to that for a low carbon steel (SAE1015), namely
TRC samples,  UTS = 350 MPa and εf = 8.5%, were more than 40%  f = 725 MPa [27]. In contrast, lower  f values are obtained for the
higher than those for the CD and GC samples, but slightly lower GC and CD samples,  f = 276 MPa for GC and  f = 385 MPa for the
than for the OCC samples ( UTS = 385 MPa and εf = 12%). The ten- CD sample. Such poor fatigue properties for the GC and CD samples
sile properties of the four cast samples demonstrate similar trends are related to their microstructural characteristics, as mentioned
to the micro-hardness, as shown in Fig. 3. The  UTS values for above. From the above experimental results, it is considered that
the OCC samples are close to those for the high quality die-cast TRC and OCC samples have excellent mechanical properties.
samples [6]. It should be pointed out that the gravity cast sam-
ple shows low tensile strength and low strain to failure, although 3.3. Microstructural analysis
some metals having low tensile strength could have high ductil-
ity. This would be caused by the presence of large brittle eutectic To understand the influence of the microstructural characteris-
structures. tics on the mechanical properties in detail, the crystal orientations
Several researchers have reported the influence of secondary were investigated. Fig. 7(a) depicts the image quality map and the
dendritic arm spacing (SDAS) on the tensile properties, and a clear crystal orientation maps of the microstructure of our cast samples.
Hall–Petch relationship was obtained in the work of Osório et al., The color level of each pixel in the crystal orientation map is defined
where a clear relationship  0.2 = 6.6d−0.5 + 40.1 was obtained for according to the deviation of the measured orientation, as indicated
the die-cast samples (ADC10) [26]. In the present work, we have in the stereographic projection. The black solid lines in the image
also examined the Hall–Petch relationships. Fig. 5 represents  0.2 quality map show grain boundaries, in which the misorientation
vs. SDAS for all the cast samples. Although there are few experi- angle is defined more than 5◦ . On the other hand, the white solid
mental data, there is a clear relationship  0.2 = 6.1d−0.5 + 48.5 with a lines show misorientation angles of less than 5◦ . Different crys-
correlation rate (R2 ) of 0.93. Due to the similar nature of the Al alloys tal orientations are found, depending on type of casting. In the GC
(ADC10 vs. ADC12), our Hall–Petch relationship is very close to that sample (Fig. 7(a)), a relatively uniform lattice orientation can be
of Osório et al. seen over a large area, where several grains seem to be united to
Fig. 6 presents the relationship between the stress amplitude make a large colony (Fig. 7(b)). It is interesting to mention that the
and fatigue life (S–N curve). Note that the arrows in this figure large misorientation angles (white line) in each grain are clearly
indicate the test specimens which did not fail within 107 cycles, observed in our GC sample, which could be a result of the high dis-
the endurance limit. It is seen that the fatigue strength of the OCC location density [28]. Furthermore, such a value of misorientation
sample is slightly higher than that for the TRC sample. The mean angle may be created by the more severe internal stress, arising
endurance limits ( end ) for the TRC and OCC samples are about from the shrinkage force. In this case, the shrinkage force will be
180 MPa and 200 MPa, respectively. On the other hand, the fatigue high, as a large amount of Al alloy (600 mm × 90 mm × 40 mm) was
strength for the GC and CD samples is almost the same;  end is used to prepare the sample.
about 100 MPa. The S–N relationships are expressed with a power In contrast, different crystal orientations can be seen in the
law dependence of the applied stress and cycle number to final CD sample, depending on the grain (Fig. 7(a)). The varying crystal
192 M. Okayasu et al. / Materials Science and Engineering A 543 (2012) 185–192

orientations in each grain are influenced by the die-casting process. GC samples. Such an excellent tensile property for TRC and OCC
The molten aluminum alloy was injected to the mould at high speed samples is caused by the fine round ˛-Al phase and tiny eutec-
and high pressure, so the melt is splashed, scattered or sprayed tic structures. As with the tensile properties, high hardness and
into the mould. It is also seen in Fig. 7(a) and (b) that no clear mis- fatigue strength are obtained for the TRC and OCC samples. With
orientation angle is detected in CD compared to GC. The different the tensile properties of the cast samples, a clear Hall–Petch
misorientation profile produces different mechanical properties, relation with  0.2 = 6.1d−0.5 + 48.5 (R2 = 0.93) is obtained. Fur-
such that a large misorientation angle produces high internal stress thermore, from the measured fatigue properties, the S–N curve
or high distortion energy, leading to the high mechanical strength. shows a power law dependence of the form  a =  f Nf −b .
Thus, the mechanical properties, such as hardness,  UTS and  f , for (3) With the EBSD analysis, lattice orientations are similarly
the GC sample are close to those for the CD samples (Figs. 3–6) observed over a large area for the GC sample, where several
even though the size of ˛-Al phase and eutectic structures for GC grains are united to make a large colony and with a high mis-
are much larger than those for CD. orientation angle in each grain. This feature produces high
Rather like a true single crystal, the crystal orientation for the mechanical properties due to the high dislocation density. The
OCC sample is almost perfectly orientated, as seen in the pole fig- CD sample shows different crystal orientation depending on
ure (Fig. 7(a)). This occurrence must be caused by the unidirectional the grain, which are produced by the injection system; molten
solidification in the cast rod. The crystal direction 1 0 0 is formed Al alloy is splashed or scattered into the mould. For the OCC
perpendicular to the casting direction of the round bar sample sample, the crystal orientation is almost regular, due to the uni-
(ND), and the crystal directions 1 1 1 and 1 1 0 can be detected directional solidification process. For the TRC sample, a slip-like
in the TD and RD directions, respectively. The uniformly oriented formation is identified in the phases side by side because of the
crystal structure in the OCC sample can lead to high ductility (εf ) high rolling force.
although the high mechanical properties are attributed to the tiny
microstructures (Fig. 4). Acknowledgement
For the TRC sample (Fig. 7(a)), the crystal orientation is seen
to be aligned, which is rather similar to that for the OCC pro- This work was supported by the grant from Sumitomo Electric
cess. However, the striped shapes of the lattice orientation at Industries CSR Foundation and Japanese Government (Ministry of
about 45◦ are identified in the phases side by side. This could be Education, Science, Sports and Culture).
more clearly observed in their pole figures. Such a striped shape
might be influenced by the slip-like formation, mechanically cre- References
ated during the rolling process. In this case, the slip would have [1] M. Okayasu, K. Kanazawa, N. Nishi, J. Jpn. Foundry Eng. Soc. 70 (1998) 779–785.
occurred mainly along (1 1 1) and 1 1 0. From this, the TRC sam- [2] R. Kimura, M. Yoshida, G. Sasaki, J. Pan, H. Fukunaga, J. Mater. Process. Technol.
ples would have high internal stress, resulting in high mechanical 130–131 (2002) 299–303.
[3] H. Iwahori, K. Tozawa, T. Asano, Y. Yamamoto, M. Nakamura, M. Hashimo, S.
properties. Venishi, J. Jpn. Inst. Light Met. 34 (1984) 525–530.
[4] K. Shinozawa, S.-M. Sun, Metall. Mater. Trans. A 28A (1997) 1441–1447.
3.4. Fractography [5] T. Haga, T. Nishiyama, S. Suzuki, J. Mater. Process. Technol. 133 (2003)
103–107.
[6] M. Okayasu, S. Yoshifuji, M. Mizuno, M. Hitomi, H. Yamazaki, Int. J. Cast Met.
Fig. 8 displays SEM images showing the fracture surfaces of the Res. 22 (2009) 374–381.
fatigue crack growth regions for the GC and OCC samples. Note that [7] D.N. Li, J.R. Luo, S.S. Wu, Z.H. Xiao, Y.W. Mao, X.J. Song, G.Z. Wu, J. Mater. Process.
Technol. 129 (2002) 431–434.
failure patterns for the GC and CD samples are similar to those of the
[8] K. Xia, G. Tausig, Mater. Sci. Eng. A A246 (1998) 1–10.
TRC and OCC samples. It is clear that different fracture modes are [9] B. Zhang, J. Cui, G. Lu, Mater. Sci. Eng. A A355 (2003) 325–330.
present. The cleavage-like fracture surfaces are the main feature in [10] H. Soda, G. Motoyasu, A. Mclean, C.K. Jen, O. Lisboa, Mater. Sci. Technol. 11
the GC sample, which implies a brittle fracture mode. On the other (1995) 1169–1173.
[11] D. Bhardwaj, H. Soda, A. McLean, Mater. Charact. 61 (2010) 882–893.
hand, a relatively rough fracture surface can be observed in the [12] Y. Wang, H.-Y. Huang, J.-X. Xie, Mater. Lett. 65 (2011) 1123–1126.
OCC samples. Such rough fracture surfaces might be created by the [13] M. Okayasu, S. Yoshie, Mater. Sci. Eng. A 527A (2010) 3120–3126.
stretched material during the loading process, giving high strength [14] M. Okayasu, S. Takasu, S. Yoshie, J. Mater. Process. Technol. 210 (2010)
1529–1535.
and high ductile materials. In addition, the rough fracture surfaces [15] Ch. Gras, M. Meredith, J.D. Hunt, J. Mater. Process. Technol. 167 (2005) 62–72.
produce more severe roughness-induced crack closure, resulting in [16] H. Chen, S.B. Kang, H. Yu, H.W. Kim, G. Min, Mater. Sci. Eng. A 492 (2008)
high fatigue strength [29,30]. 317–326.
[17] S.X. Zhou, J. Zhong, D. Mao, P. Funke, J. Mater. Process. Technol. 134 (2003)
363–373.
4. Conclusions [18] A. Ohno, J. Met. 38 (1985) 14–16.
[19] M. Okayasu, S. Takasu, M. Mizuno, J. Mater. Sci. 45 (2010) 1220–1226.
[20] M. Okayasu, K. Sakai, S. Takasu, Int. J. Cast Met. Res. 24 (2008) 286–298.
The material properties of cast aluminum alloy ADC12, pro- [21] L.Y. Zhang, Y.H. Jiang, Z. Ma, S.F. Shan, Y.Z. Jia, C.Z. Fan, W.K. Wang, J. Mater.
duced by the GC, CD, TRC and OCC processes, have been studied Process. Technol. 207 (2008) 107–111.
[22] A.M. Samuel, F.H. Samuel, J. Mater. Sci. 30 (1995) 1698–1708.
experimentally. The following conclusions can be drawn. [23] N. Saheb, T. Laoui, A.R. Daud, M. Harun, S. Radiman, R. Yahaya, Wear 249 (2001)
656–662.
[24] L.A. Dobrzański, W. Borek, R. Maniara, J. Achievements Mater. Manuf. Eng. 18
(1) Different microstructural characteristics are obtained. For (2006) 211–214.
example, the microstructure for GC and CD samples consists [25] L.A. Dobrzański, R. Maniara, J. Sokłowski, W. Kasprzak, J. Mater. Process. Tech-
mainly of the ˛-Al phase, eutectic CuAl2 , eutectic Si and Fe. In nol. 191 (2007) 317–320.
[26] W.R. Osório, P.R. Goulart, G.A. Santos, C.M. Neto, A. Garcia, Metall. Mater. Trans.
contrast, fine round ˛-Al phase and tiny silicon and iron eutec- A 37 (2006) 2525–2538.
tic structures are observed for the TRC and OCC samples. The [27] R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering Materials,
˛-Al phases for the TRC sample are strained severely due to the 4th ed., John Wiley & Sons, Inc., New York, 1996, pp. 556–570.
[28] P.D. Littlewood, T.B. Britton, A.J. Wilkinson, Acta Mater. 59 (2011) 6489–6500.
rolling process.
[29] M.R. Parry, S. Syngellakis, I. Sinclair, Mater. Sci. Eng. A291 (2000) 224–234.
(2) Ultimate tensile stress and strain to failure for the TRC and OCC [30] S.H. Wang, C. Müller, H.E. Exner, Metall. Mater. Trans. A 29A (1998)
samples are more than 40% higher than those for the CD and 1933–1939.

You might also like