Você está na página 1de 8

Materials Science and Engineering A 472 (2008) 107114

Microporosity control and thermal-fatigue resistance


of A319 aluminum foundry alloy
H. Arami a,b , R. Khalifehzadeh a,b , M. Akbari a , F. Khomamizadeh a,
a Department of Materials Science and Engineering, Sharif University of Technology, P.O. Box 11365-9466, Azadi Ave., Tehran, Iran
b Materials and Energy Research Center, P.O. Box 14155-4777, Tehran, Iran

Received 6 May 2006; received in revised form 2 March 2007; accepted 10 March 2007

Abstract
The objective of this work was to gain a quantitative and qualitative understanding of the degree to which porosity influences thermal-fatigue
performance and other mechanical properties of A319 aluminum casting alloy over ranges of commercial interest. The number of cycles to
failure for each test was recorded. An increase in porosity content caused degradation in thermal-fatigue life and other mechanical properties. The
fractographic examinations identified the pores and some intermetallics as the key microstructural features which promote damage and thermal-
fatigue crack initiation sites in the specimens. Crack initiation and propagation is expected to occur sooner in regions of higher defects such as pores
and large intermetallics. Progressive cyclic plastic deformation was also observed in constrained thermal-fatigue specimens. In the absence of other
defects, thermal-fatigue crack initiation is dominated by accumulation of the local plastic deformation in the -Al in constrained thermal-fatigue
fracture surfaces.
2007 Elsevier B.V. All rights reserved.

Keywords: A319 aluminum alloy; Thermal-fatigue; Microporosity; Intermetallic

1. Introduction must battle. Thus, it is very important to understand its effects


on the properties of castings. Wang et al. [4] investigated the
Among cast aluminum alloys, 319 ranks as one of the com- influence of casting defects on the room temperature fatigue per-
mercially important aluminiumsilicon (AlSi) alloys mostly formance of a Sr-modified A356-T6 casting alloy and reported
used in fabrication of cylinder heads and engine blocks, on that compared with oxide films, porosity is more detrimental to
account of its excellent casting characteristics, good mechan- fatigue life. It is interesting to note that the porosity increases
ical properties, low cost component production, crucial energy with decreasing the degassing time and depending on the appli-
efficiency and concomitant environmental benefits [1]. In these cation, the size and/or quantity of porosity to be allowed may
components processing, design and application, it is critical vary considerably. High performance aerospace castings will
to comprehend the effect of microstructures on properties [2]. generally allow none, while commercial automotive cylinder
Rakopoulos et al. [3] reported a combined experimental and the- heads will have critical requirements in areas like the combustion
oretical investigation of the phenomenon of short term response face, and noncritical automotive parts [5]. Studies have demon-
(cyclic) temperature transients in the combustion chamber walls strated that whenever present at or near the specimen surface,
of a reciprocating internal combustion engine and demonstrated these defects can cause rapid crack initiation and crack growth
the influence of the engine operating conditions and wall mate- will dominate the fatigue behavior. Seniw et al. [6] reported that
rial properties on its fatigue resistance. the microstructural heterogeneity of aluminum A356 permanent
The casting process inevitably introduces defects, such as mold castings impacts fatigue performance differently depend-
porosity, inclusions and entrapped oxide films in the solidified ing on location in the castings. Heuler et al. [7] and Dabayeh et al.
metal. Porosity is the most common casting defect that foundries [8] reported that surface and near surface defects initiate fatigue
cracks sooner than defects embedded in the specimen. Buffiere
et al. [9] studied the microstructure and fatigue properties of
Corresponding author. Tel.: +98 21 66165210; fax: +98 21 66005717. three model AS7G03 cast aluminum alloys containing artificial
E-mail address: khomami@sharif.edu (F. Khomamizadeh). pores. They concluded that both the average number of cycles to

0921-5093/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.03.031
108 H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114

failure and the lifetime scatter depend on the pore content espe-
cially at high stress level. They identified the mechanism leading
to the initiation of a crack from a pore and concluded that crack
propagation at high stress level appears to be quite insensitive to
microstructural barriers. At low stresses, however, short cracks
are often observed to be stopped at grain boundaries and the
fatigue life is no longer predicted by a simple propagation law.
Gao et al. [10] investigated six different, industrially important,
casting situations containing an array of typical microstructures
and defects. They identified porosity, secondary dendrite arm
spacing, Al-matrix, Si-particles and Fe-rich intermetallics as the
major factors affecting the alloys resistance to fatigue.
Even though a large number of such studies have been, and
are continuously being, reported in the literature, very few deal
with the fact that their operating environment is structurally
demanding and a deeper understanding of their thermal-fatigue
performance is required for engine design. It is therefore impor-
tant to investigate the interaction between the microstructural
features and the thermal-fatigue damage evolution of this par-
ticular alloy. Here, we examine a particular microstructure, i.e. Fig. 1. Constrained (a) and flat (b) thermal-fatigue specimens design.
porosity, known to be related to mechanical properties in alu-
minum castings such as A319, and attempt to investigate its
effects on thermal-fatigue performance of the alloy, using a Advantage Image Analysis system. Tensile tests were performed
quantitative approach. at room temperature by Instron 6027 machine, in accordance
with ASTM B557M (1 mm/min strain rate). The yield stress
2. Experimental methods was based on a 0.002 plastic strain offset. The hardness was
measured using an ERNST Brinell hardness tester (2.5 mm ball
The material used in this work was A319 cast alloy, the chem- 31.25 kg load).
ical composition being given in Table 1. Alloys were melted The cast bar was cut up into two types of constrained cylin-
in a silicon crucible of 8 kg capacity using an electrical resis- drical and flat thermal-fatigue samples which are presented by
tance furnace. The melting temperature was kept at 735 5 C. Fig. 1. The constrained thermal-fatigue test is a test method
Degassing of the melt was carried out by an argon rotary originally designed to evaluate materials exposed to sustained
degassing installation (15 min). The modification of the melt was thermal stresses, such as exhaust manifolds and cylinder heads.
made by Al10%Sr at 13th minute of degassing process to pre- The test measures the influence of constrained (axial) ther-
vent strontium fading during melt degassing. The casting process mal expansion and contraction on thermal-fatigue life. It has a
consisted of pouring the molten metal into a 200 C permanent cylindrical geometry measuring 140 mm long with button-head
mould. Care was taken to minimize turbulence during pouring shoulders of 20 mm in diameter by 14 mm long. The gauge sec-
to avoid gas pickup and entrainment of inclusions. After homog- tion is 11mm in diameter by 112 mm long. A 4.8 mm diameter
enization and quench, an aging treatment leading to a peak-aged hole is drilled and reamed through the central, longitudinal axis
state labeled T6 has been carried out, providing the maximum of the specimen to accommodate the control thermocouple. The
hardness of the aluminum matrix. The expected porosity was test equipment consisted of a rigid test frame with an opening
revised by a density test, performed on as-cast specimens for of about 15 mm, a gas flame as the heater and a waterair mix-
three times. A Startorius-CP324S densitometer was used to ture cooling system. The specimens were mounted in a frame
determine the density of the end products by Archimedes prin- in order to prevent its axial expansion and contraction dur-
ciple. For optical image analysis of specimens and hardness ing heating and cooling. The specimen was preheated to about
measurements, some random samples were excised from the 100 C and allowed to stabilize at temperature for a minimum
center of the castings. The surfaces of the samples were ground of 15 min, then locked in the fixture. After appropriate locking,
with 1200 grit silicon paper. The microstructures of the materials the sample was alternatively heated to the selected peak tem-
were investigated with an Olympus-PME3 optical microscope perature (250 C) for about 100 10 s, and then cooled rapidly
and their porosity contents were measured with an Omnimet to a lower temperature (40 5 C). Repetitive thermal cycling
results in thermal-fatigue damage and, ultimately, fatigue crack
propagation.
Table 1 Surfaces of the flat thermal-fatigue specimens were ground
Chemical composition of the studied A319 alloy (wt%) down to 4000 grit, followed by mechanical polishing with 1 m
Al Si Cu Fe Mn Zn Mg Others diamond. The final polishing was carried out using a colloidal
89.93 6.02 3.041 0.694 0.2007 0.0432 0.0155 0.0556
silica suspension. The principles of heating and cooling cycles in
these specimens experiments were the same as the constrained
H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114 109

thermal-fatigue tests. But there was not any limitation for sam-
ples expansion and contraction during their heating and cooling.
The notches of flat samples were machined in the middle of
each sample, using electro-discharge machining (EDM) for pre-
cision cutting. A gas flame and circulating 25 C water were
used for specimens heating and cooling. An alumelcrumel
thermocouple was used to control the specimen temperature
during the test. Each sample was alternatively heated to the
selected peak temperature (230 15 C) for about 55 s, and then
abruptly cooled to a lower temperature (23 C). The repetitive
thermal cycling results in thermal-fatigue crack propagation in
the regions near to the notch root. In order to investigate the
crack initiation and propagation mechanisms, the surface of
the specimens was characterized by optical microscope (OM)
during the tests. The examinations were carried out after 200
cycles, and then every 50 cycles. Such tests were interrupted
when a main crack had attained a size of about 1200 m, where
it became quite insensitive to microstructural effects. Fracture
surfaces were examined in a SEM-Philips-XL40. For scanning
electron microscopy observations of small-crack initiation and
growth, flat specimens were polished with 1 m diamond paste.
Three specimens were tested for each condition of both con-
strained and flat thermal-fatigue experiments and the average
results were reported for each.

3. Results and discussions

Metallographic examinations have shown that the basic


microstructural constituents of cast A319 comprise a primary
Al-matrix together with an Al/Si eutectic located between the
secondary dendrite arms [1]. In addition, pores in as-cast mate-
rials are classified into two types. One is the open pore that
connects to the surface of the material. The other is the closed
pore that is isolated inside material [11]. Fig. 2 shows opti-
cal micrographs of as-cast specimens after different degassing
intervals. As it is shown typically, intermetallic phases, pores,
Al-matrix and Si-particles are the major features in the cast
alloy. The average secondary dendrite arm spacing (SADS) of
microstructures was about 45 m. Images show the porosity
content of specimens as a function of degassing process time.
It can be seen that the number of open pores and their sizes
increase with decreasing the degassing process interval. Evolv-
ing gasses which are trapped at the solidliquid interface and
dendrite arms during the cooling and solidification, form gas Fig. 2. OM images of typical microstructural constituents in cast specimens
pores which weaken mechanical properties of the cast metal [1]. after degassing for: 25 min (a); 10 min (b); 0 min (c).
However, as it is shown in Fig. 3 degassing process decreases
porosity content in the cast samples, because the H2 -content of the specimen hardness. It seems that an increase in porosity con-
the molten metal decreases [12]. There is a noticeable similar- tent of the cast specimens causes degradation in their general
ity between the data obtained from image analysis method and mechanical properties.
archimedes principle. But of particular interest to this work was to gain an
Tensile tests have been performed on as-cast and T6 speci- understanding of the degree to which porosity influences
mens for three times and their results are reported in Fig. 4. As it thermal-fatigue performance of A319 aluminum casting alloy
is shown, increasing the porosity content in the cast specimens over ranges of commercial interest. The measured thermal-
decreases yield stress of the samples progressively. Fig. 5 shows fatigue lifetimes as a function of porosity content are shown
the average values of hardness (Brinell) as a function of poros- in Fig. 6. The experiments have been done before and after T6
ity content. Increasing the porosity in specimens decreases the heat treatment. As it is shown, the thermal-fatigue resistance is
matrix strength against the indentor pressure, and this decreases proportional to the porosity content and increasing the poros-
110 H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114

Fig. 3. Porosity content (Vol.%) of cast specimens after different periods of


degassing process measured by Archimedes principle and image analysis sys-
tem.

Fig. 6. Cycles to failure in constrained (a) and flat (b) as-cast and T6 thermal-
fatigue specimens as a function of porosity content.
Fig. 4. Tensile yield strength of as-cast and T6 specimens as a function of
porosity content.
raisers in front of the notch root. If an external force is applied,
there is a stress concentration around the pores. In the thermal-
ity content, decreases the thermal-fatigue life. The surfaces of fatigue condition, the stress concentration around the pores may
the flat specimens were investigated during the experiments, cause crack propagation. Note that pores in the sizes observed
using optical microscope. Fig. 7 shows the length of the cracks in this work, and in their very tortuous shape, did not cause
in thermal-fatigue flat specimens (as-cast and T6 specimens) crack blunting. Buffiere et al. [9] identified the mechanism lead-
as a function of the number of cycles and the porosity con- ing to the initiation of a crack from a pore. They reported that
tent. Existence of pores in front of a crack reduces the matrix almost all the pores are intergranular and the number of grain
resistance against crack propagation and increases the thermal- boundaries starting from the surface of a pore increases with the
fatigue crack propagation rate, especially when they act as stress pore size. Besides, the convex parts of the pores are in all cases
associated with Si eutectic particles, and also with at least one
grain boundary. This typical arrangement of pores, Si-particles
and grain boundaries, has crucial importance with respect to the
crack initiation mechanisms. Notice that in the problem under
study, we have supposed that the cavity occurs at the notch root.
In addition, once the porosity content which directly controls the
fatigue life in our tests is properly accounted for, other parame-
ters, such as stress amplitude, largest pore size, DAS, modifier
content and grain refinement lose their significance. Of course
this is not to say that these are unimportant. Indeed, they have
been chosen as constant parameters which do not influence the
results.
Regarding to Figs. 46, it can be obviously seen that after
solution heat treatment and artificial aging mechanical proper-
Fig. 5. Average hardness of the as-cast and T6 specimens as a function of ties of the cast specimens have been substantially improved by
porosity content. precipitation heat treatment. The most important effect of the
H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114 111

Fig. 7. Length of the crack in as-cast (a) and T6 heat treated (b) flat thermal-
fatigue specimens as a function of the number of cycles and the porosity content.

formation of the Al2 Cu precipitates as the second phase is that


the aluminum matrix is hardened. This phenomenon improves Fig. 8. SEM photographs showing crack initiation in flat (a) and cylindrical (b)
thermal-fatigue specimens cast without degassing.
mechanical properties of the alloy considerably. As might be
expected, formation of reinforcing Al2 Cu coherent precipitates
in the peak-aged (T6) specimens leads to a reduction in thermal- Fig. 9 shows typical constrained thermal-fatigue fracture sur-
fatigue crack propagation rates. This results from the reduction faces, revealing details in the vicinity of gas and shrinkage
in cyclic strain within the matrix and transfer of cyclic loads to pores. It is reported that presence of the pores in the regions
the precipitates. Furthermore, as cracking proceeds within the near the notch root, may cause a larger strain concentration
matrix, unbroken precipitates remain behind the advancing crack because of their somehow elongated shape and sharp tips. These
front and restrict crack opening. This crack-tip shielding mech- microstructural features give rise to an inhomogeneous distribu-
anism, leads to vastly reduced fatigue crack propagation rates tion of stresses and strains at the micro-scale [9]. Consequently,
and as a result, crack initiation and propagation resistance in in specific regions, such as the tip of the pores, the local micro-
the age-hardened material increases [13]. However, holding, or stresses are higher than the average and may exceed the local
aging, the specimens for too long a period at a given temperature yield stress. The magnitude of this micro-plastic deformation
causes them to lose their hardness, since overaging may occur. is dependent on both the far-field stress level and the nature
Caton et al. [14] reported that while the peak-aged (T6) condi- of the inhomogeneous features. Generally, under a load less
tion has a significantly higher tensile strength, the growth rates than the nominal yield stress, the average magnitude of the
of small cracks were only slightly slower in these specimens micro-plastic strain is very small. However, under cyclic loading,
when compared to the over-aged (T7) condition for equivalent with increasing number of cycles the micro-plastic deforma-
stress levels and manufacturing parameters. tion can be accumulated to a critical local value which is large
The thermal-fatigue crack initiation sites were identified by enough to initiate fatigue cracks. This accumulated plastic strain
SEM fractography. In samples cast without degassing, cracks is significant in the development of thermal-fatigue damage in
initiated from the notch tip and propagated through pores located constrained specimens, particularly in the regions near the notch
at, or close to, the notch root region. Fig. 8 shows typical SEM surfaces. Since the samples were heated and cooled from the
photographs of specimens cast without degassing, after thermal- surface, defects which are near the surface induce higher stress
fatigue crack initiation. concentration than defects in the interior of the specimen. There-
112 H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114

Fig. 9. Typical constrained thermal-fatigue fracture surfaces with an initiating


gas (a) and shrinkage (b) pore.

Fig. 10. Typical SEM image and EDAX analysis from a thermal-shock crack
fore, the pores at or near the specimen surface induce much
propagated through the pores (a) and developed by de-bonding and cracking of
higher stress/strain concentration than other pores in the interior large Cu rich (a) and CuMg rich (b) intermetallics.
and the plastic deformation is higher at the surface and local
surface strains can exceed the average strain of the specimen
by orders of magnitude. Panda et al. [15] analyzed the effect of intermetallics are weakly bonded to the matrix and form inco-
temperature distribution on hot and cold faces of thermal-fatigue herent interfaces with the matrix, crack deflection occurs as the
specimens and showed that the thermal stress is also sensitive to crack moves the matrix around the particles. This crack deflec-
thickness of the sample. Sherman and Schlumm [16] explained tion lead to enhanced fatigue crack propagation and it can be
this phenomenon due to the increase in compressive stress on the concluded that the probability of crack initiation and propaga-
surface with decrease in sample thickness and reported that the tion is higher from these intermetallics. Not only do the large
overall tensile stress decreases on the quench surface. Sjostrom incoherent intermetallics nucleate cracks, but also they aid in
and Bergstrom [17] estimated the surface mechanical strains their propagation through the matrix and even their jumping
during thermal cycling from the laser speckle measurements. from one to the next.
Thermal-fatigue cracks passing across the pores in the regions If the specimen has preferred microstructural orientations
near the notch root can be seen in Fig. 10a. Crack propagation in which neighboring grains and dendrites have slip planes
will be faster in regions of higher porosity since they are free with different orientations, it is possible for slip-band cracks
spaces (voids) without strength. Images also show propagation to extend across grain boundaries and propagate through dif-
of the thermal-fatigue crack through some large intermetallics ferent preferred directions. Fig. 11 shows the propagation of a
formed during the solidification of the molten alloy. Regarding multidirectional thermal-fatigue crack in the regions close to the
to EDAX compositional measurements, it seems that the white notch root.
regions (Fig. 10a) and the spherical particle (Fig. 10b), which The fractograph in Fig. 12a is related to the thermal-fatigue
can be seen in the vicinity of the cracks, are Cu rich and CuMg fracture surface of a constrained specimen, which was pulled
rich intermetallics, respectively. The prevailing thermal-fatigue to failure. The observations of the fracture surface show typ-
fracture mechanisms are strongly dependent on the respective ical plastic deformed pattern. In most of the cases the cracks
properties of the matrix, interface and precipitates. When strong were found to originate from the Al-matrix near the specimen
H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114 113

micro-cracks were linked by the localized plastic deformation


in the form of shear bands to form the dominant small crack
[18]. The linkage of micro-cracks through the Al-matrix created
the faceted surface observed in this image [19]. This progres-
sive damage accumulation led to final fracture of the constrained
thermal-fatigue specimens. Diaz et al. [20] used a non-contact
digital image measurement system to study the fatigue damage
accumulation around a notch in 304 stainless steel and showed
that during the first part of the crack growth stage, the externally
given work not only impels the growth of the evaluated crack and
its local plastic zone but also generates plastic damage around
the notch, away from the crack influence. They reported that
damage generation around the notch is cyclic and a low cycle
number is enough to generate a new plastic zone. Also, the effect
Fig. 11. Propagation of a multidirectional thermal-fatigue crack in the regions
of constraint on ductile fracture initiation from a notch tip under
close to the notch root in cylindrical thermal-fatigue specimens. mode I and mixed mode (involving modes I and II) loading is
investigated by Roy et al. [21].
surface. The localized and concentrated plastic deformation in Cylindrical thermal-fatigue specimens were under con-
the matrix could be sufficient to form a slip band across it and strained (axial) thermal expansion and contraction. This
initiate a fatigue crack. Indeed, in constrained thermal-fatigue phenomenon, which was eliminated in flat thermal-fatigue
specimens there was significant plastic deformation of the - experiments, caused progressive plastic deformation in the
Al leading to formation of microcracks in the regions near the vicinity of the cracks surfaces in constrained thermal-fatigue
notch root. The micro-cracks initially developed, due to the local specimens. In contrast, as Fig. 12b shows, in flat specimens, a
accumulation of plastic deformation within the cells in the pref- brittle fracture was observed which is directly related to their
erentially oriented grains and intersecting the surface. These stress-free mode during the experiments.

4. Conclusion

The results of this work highlight the importance of poros-


ity control in manufacturing of thermal-fatigue critical structural
components. An increase of porosity content causes degradation
in thermal-fatigue resistance and other mechanical properties of
cast 319 aluminum alloys. Under repeated loadings thermal-
fatigue cracks propagate through casting defects like pores and
large incoherent intermetallics in regions near the notch root,
where there are stress concentrations. T6 heat treatment substan-
tially improves the mechanical properties of the cast specimens.
Crack-tip shielding mechanism, leads to vastly reduced thermal-
fatigue crack propagation rates.

References

[1] J.R. Davis, Aluminium and Aluminum Alloys, ASM Specialty Handbook,
ASM International, Metals Park, OH, 1993.
[2] A. Hetke, R.B. Gundlach, AFS Trans. 137 (1994) 376380.
[3] C.D. Rakopoulos, D.C. Rakopoulos, G.C. Mavropoulos, E.G. Giakoumis,
Appl. Therm. Eng. 24 (2004) 679702.
[4] Q.G. Wang, D. Apelian, D.A. Lados, J. Light Met. 1 (2001) 7384.
[5] M. Nakai, T. Eto, Mater. Sci. Eng. A285 (2000) 6268.
[6] Mark E. Seniw, James G. Conley, Morris E. Fine, Mater. Sci. Eng. A285
(2000) 4348.
[7] P. Heuler, C. Berger, J. Motz, Fatigue Fract. Eng. Mater. Struct. 16 (1992)
115136.
[8] A.A. Dabayeh, A.J. Berube, T.H. Topper, Int. J. Fatigue 20 (1998) 517530.
[9] J.-Y. Buffiere, S. Savelli, P.H. Jouneau, E. Maire, R. Fougeres, Mater. Sci.
Eng. A316 (2001) 115126.
[10] Y.X. Gao, J.Z. Yi, P.D. Lee, T.C. Lindley, Acta Mater. 52 (2004) 5435
5449.
Fig. 12. Typical thermal-fatigue fracture surfaces and plastic deformation of the [11] S. Shivkumar, D. Apelian, J. Zou, AFS Trans. 98 (1990) 897904.
-Al in cylindrical (a) and flat (b) specimens. [12] R.C. Atwood, S. Sridhar, W. Zhang, Acta Mater. 98 (2000) 405417.
114 H. Arami et al. / Materials Science and Engineering A 472 (2008) 107114

[13] R.E. Reed-Hill, R. Abbaschian, Physical Metallurgy Principles, third ed., [18] J.Z. Yi, Y.X. Gao, P.D. Lee, T.C. Lindley, Mater. Sci. Eng. A386 (2004)
PWS Publishing Company, Inc., Boston, USA, 1994. 396407.
[14] M.J. Caton, J.W. Jones, J.E. Allison, Mater. Sci. Eng. A314 (2001) 8185. [19] Q.G. Wang, D. Apelian, D.A. Lados, J. Light Met. 1 (2001) 8597.
[15] P.K. Panda, T.S. Kannan, J. Dubois, C. Olagnon, G. Fantozzi, Sci. Technol. [20] F.V. Diaz, A.F. Armas, G.H. Kaufmann, G.E. Galizzi, Opt. Lasers Eng. 41
Adv. Mater. 3 (2002) 327334. (2004) 477487.
[16] D. Sherman, D. Schlumm, Scripta Mater. 42 (2000) 819825. [21] Y. Arun Roy, R. Narasimhan, P.R. Arora, Acta Mater. 47 (March (5)) (1999)
[17] Johnny Sjostrom, Jens Bergstrom, J. Mater. Process. Technol. 153154 15871596.
(2004) 10891096.

Você também pode gostar