You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260060167

Solidification Structure of Aluminum Alloys

Chapter · January 2004


DOI: 10.13140/RG.2.1.2892.0169

CITATIONS READS

3 2,788

2 authors:

Doru M. Stefanescu Roxana Ruxanda


The Ohio State University Emerson
365 PUBLICATIONS   4,320 CITATIONS    24 PUBLICATIONS   302 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nucleation and growth of spheroidal graphite View project

All content following this page was uploaded by Doru M. Stefanescu on 20 February 2014.

The user has requested enhancement of the downloaded file.


Name /bam_asmint_106420/6044_002d5/Mp_1 07/21/2004 02:27PM Plate # 0 pg 1 #

Solidification Structures
of Aluminum Alloys
Doru M. Stefanescu and Roxana Ruxanda, The University of Alabama

ALUMINUM-BASE ALLOYS constitute a system is similar to that in Fig. 16(b) of the ar- microstructure. This is demonstrated in Fig. 3
group of cast materials second only to ferrous ticle “Fundamentals of Solidification” in this and 4 for a nominal Al-4.5Cu-0.25Mg alloy with
castings in tonnage. Aluminum alloys have Volume, representing partial solid solubility with a small amount of silver (0.7%) and titanium,
gained increasing markets in the automotive and a eutectic reaction. The eutectic invariant is at respectively. Titanium addition considerably de-
aircraft industry. New alloys are still under de- 11.7% Si and 577 ⬚C (1070 ⬚F). Typical exam- creases the grain size.
velopment, since there are still wide possibilities ples of hypoeutectic, eutectic, and hypereutectic Aluminum-Magnesium Alloys. The corner
of improving their properties by using simple, aluminum-silicon commercial alloys are given in of the phase diagram of practical interest is simi-
affordable techniques, such as alloying and heat Fig. 1 (Ref 2). lar to that of aluminum-copper alloys, with the
treating. A wide range of materials can be added Binary eutectic or hypoeutectic aluminum-sil- maximum solubility of magnesium in aluminum
to aluminum (Ref 1): zinc, magnesium, copper, icon alloys are characterized by good castability at 17.4% Al. Some typical examples of alumi-
silicon, iron, lithium, manganese, nickel, silver, and corrosion resistance. However, the alumi- num-magnesium commercial alloys are given in
tin, and titanium. The solid solubility of these num-silicon alloys are seldom binary. Strength- Fig. 5.
elements in aluminum varies considerably. Some ening of aluminum-silicon alloys is achieved by Other Multicomponent Aluminum Alloys.
elements are used as solid-solution strengthen- adding small amounts of other elements, such as Investigations of higher-order multicomponent
ers, others form desirable intermetallic com- copper, magnesium, or iron. Higher iron con- alloys solidification are rather limited. Sophisti-
pounds. The most common alloy systems are tents can promote the formation of brittle plates cated methods are used to identify phases and
aluminum-silicon, aluminum-copper, and alu- of AlFeSi or other complex intermetallics in the their orientation. An example of the microstruc-
minum-magnesium. Some typical solidification presence of manganese. This may have a nega- ture of the Al 7050 alloy containing 11 elements
microstructures for these alloying systems are tive influence on the mechanical properties. is given in Fig. 6 (Ref 3). The dark areas are ␣-
given in the following sections. Aluminum-Copper Alloys. The equilibrium aluminum phase and the bright areas are quater-
diagram of the binary aluminum-copper system nary sigma (r) phase (Al,Cu,Zn)2Mg, ternary S
is given in Fig. 2. These alloys are typical solid- phase (Al2MgCu), or/and binary h phase
Basic Microstructures solution alloys generally containing 4 to 6% Cu (Al2Cu). These phases were identified using op-
of Aluminum-Base Alloys and some magnesium. After a solution treat- tical metallography, scanning electron micros-
ment, CuAl2 particles precipitate from the copy (SEM), x-ray map, and electron probe mi-
Aluminum-Silicon Alloys. The equilibrium quenched alloy, as seen in Fig. 2. The level of croanalysis (EPMA) techniques. The dendrite
phase diagram of the binary aluminum-silicon impurities can affect both the macrostructure and arms are separated by sigma, S, h, and/or eutec-

Fig. 1 Typical microstructures of hypoeutectic, eutectic, and hypereutectic aluminum-silicon commercial alloys. (a) Hypoeutectic aluminum-silicon alloy (Al-5.7Si, alloy type A319).
Fan-shaped Al51-(MnFe)3-Si2 phase growing in competition with the ␣-aluminum phase, silicon crystals, Al2Cu, and areas with complex eutectics. Etchant: 0.5% HF. Original
magnification 110⳯. (b) Eutectic aluminum-silicon alloy (Al-11.9Si, alloy type A339). ␣-aluminum dendrites, primary silicon particles, and areas with complex eutectics. Etchant: 0.5%
HF. Original magnification 110⳯. (c) Hypereutectic aluminum-silicon alloy (Al-15Si, alloy type A390). Large primary silicon particles, eutectic silicon crystals, and Al2Cu phase in a
matrix of ␣-aluminum phase. Etchant: 0.5% HF. Original magnification 110⳯
Name /bam_asmint_106420/6044_002d5/Mp_2 07/21/2004 02:27PM Plate # 0 pg 2 #
2 / Metallurgy and Microstructure

ticlike network, or in touch with each other. The visible especially near the primary stem and the steep temperature gradient exist during solidifi-
primary stem in Fig. 6 is located at the junction root of the tertiary arms. cation, the entire structure will remain columnar
of two secondary arms. The slightly darker gray (Fig. 7). However, in normal conditions, some
scales at the center of the dendrites show lower arms of the floating dendrites can be detached by
solute contents in the ␣ phase, indicating the Grain Structure the convective motion of the melt, thus deter-
main growth directions of the dendrite arms. mining the formation of an equiaxed structure in
Small, isolated intradendritic droplets are also Grain size is a readily observed feature of alu- the middle of the ingot (Fig. 8).
minum alloy ingots and castings. For solid-so- Grain size may be controlled by such me-
lution type alloys, mechanical properties are chanical methods as vibration, stirring, and con-
highly dependent on the primary grain size. A trol of metal flow, which provide nuclei by de-
uniform, fine grain size is sought in most in- tachment of dendrite arms (Ref 5).
stances to obtain optimal properties in the Grain-refining additions may also be used to
wrought product. Accordingly, for solid-solution change the structure from coarse and nonuniform
type alloys the size of the primary grains result- (Fig. 9a) to a fine, uniform one (Fig. 9b). The
ing from solidification is reduced by the use of grain-refining inoculants commonly used in the
grain refinement. aluminum industry are master alloys containing
The properties of cast alloys containing large titanium or titanium plus boron (Ref 6–9).
amounts of eutectic, such as the aluminum-sili- The aluminum alloy ingots cast without grain
con alloys, depend more on the eutectic mor- refiner often exhibit a fan-shaped columnar
phology and the dendrite arm spacing than on structure, referred to as “feather crystals” (Ref
the grain size. Therefore, the modification of the 4). This structure, illustrated in Fig. 10, may be
brittle silicon phase from the eutectic is primarily found in low- and high-solute alloys. It is most
used when processing aluminum-silicon alloys. likely to develop when there is a steep thermal
The primary solidification grains in aluminum gradient ahead of the solidifying interface (Ref
alloys ingots have a very pronounced columnar 5) or an inadequate addition of grain refiner (Ref
structure, directed from the mold interface to- 10). At higher magnification, the feather crystals
ward the core (Ref 4). If little turbulence and consist of twinned columnar grains (Fig. 11).
Fig. 2 A portion of aluminum-copper equilibrium dia-
In contrast to aluminum ingot alloys, the foun-
gram
dry eutectic alloys exhibit a very regular, mostly

Fig. 4 Micrographs of aluminum-copper alloy with a


Fig. 3 Micrographs of aluminum-copper alloy with a small amount of titanium, having a higher level
small amount of silver, having a low level of im- of impurities (alloy type A206 with 4.36% Cu). (a) Macros-
purities (alloy A201 with 4.10% Cu). Macrostructure show- tructure showing grains and dendrites. Etchant: 1 mL HBF4 Fig. 5 Typical examples of aluminum-magnesium com-
ing grains and dendrites. Etchant: 1 mL HBF4 1 mL HF, 24 1 mL HF, 24 mL C2H5OH, 74 mL H2O electrolyte. Original mercial alloys. (a) Microstructure showing Al3Fe
mL C2H5OH, 74 mL H2O electrolyte. Original magnifica- magnification 56⳯, observed under polarized light. See (gray) and Mg2Si (black) in ␣-aluminum solid-solution ma-
tion 56⳯, observed under polarized light. See Fig. 2d5- Fig. 2d5-4a on page XXX of the article “Selected Color Im- trix (alloy type A518 with 7.6% Mg). Etchant: 0.5% HF.
3(a) on page XXX of the article “Selected Color Images” for ages” for color version. (b) Microstructure showing Original magnification 560⳯. (b) Microstructure showing
color version. (b) Microstructure showing Al7FeCu2 (nee- Al7FeCu2 (needles) and Al2Cu in ␣-aluminum solid-solu- ternary eutectic and ␣-aluminum solid-solution dendrites
dles) in ␣-aluminum solid-solution matrix. Etchant: 0.5% tion matrix. Etchant: 0.5%HF. Original magnification matrix (alloy type A512 with 4Mg-1.8Si). Etchant: 0.5% HF.
HF. Original magnification 560⳯ 560⳯ Original magnification 560⳯
Name /bam_asmint_106420/6044_002d5/Mp_3 07/21/2004 02:27PM Plate # 0 pg 3 #
Solidification Structures of Aluminum Alloys / 3

equiaxed structure, as shown in Fig. 12. If prop- In unmodified faceted silicon crystals, growth ing diffraction (EBSD) in directionally solidified
erly etched, aluminum-silicon eutectic alloys re- is favored in certain crystallographic directions. Al-7Si and A356 alloys, unmodified and stron-
veal both the equiaxed grains and the dendrites Some twin planes that form “re-entrant edges” tium modified (Ref 15). An example of an EBSD
within each grain. Different etching techniques are particularly effective in promoting growth map in relationship with the microstructure is
(Ref 11) can be used to outline the grains, as (see Fig. 19). Some theories explain the mor- given in Fig. 22. Colors indicate different crys-
shown in Fig. 13 and 14. phological change by the disturbances in the tallographic orientations on the EBSD map.
growth step of the silicon crystals induced by the Nonindexed points are black. The relationship
adsorption of the modifying elements, which between the crystallographic orientation parallel
Eutectic Microstructure causes frequent twinning to occur (see Fig. 20) to the sample surface normal and the colors in
of Aluminum-Silicon Alloys (Ref 12). Examples of the effect of modification the EBSD maps is given at the right. It was dem-
on the microstructure of some alloys are given onstrated that the eutectic aluminum has mainly
The microstructure of the eutectic or near-eu- in Fig. 21 (Ref 13, 14). the same crystallographic orientation as the den-
tectic alloys consists of acicular or lamellar eu- New methods of investigation, capable of de- drites in the unmodified alloys and the stron-
tectic silicon (faceted crystals) dispersed termining the grain orientation, have emerged. tium-modified Al-7Si alloy. A planar eutectic
throughout the aluminum matrix. Examples are The microstructure of the eutectic growth inter- growth front was observed in the Al-7Si alloy
given in Fig. 15 and 16. The addition of various face was investigated with electron backscatter- alloys; a more complex eutectic grain structure
modifiers (sodium, calcium, strontium) in small was found in the strontium-modified A356 alloy.
amounts changes the morphology of the eutectic
phase from acicular to fibrous (nonfaceted), as
shown in Fig. 17 and 18. Dendritic Microstructure

An example of a microstructure containing


typical aluminum dendrites is shown in Fig. 23.
The dendrites, revealed by deep etching the mi-
crostructure with modified Poulton reagent, were
enhanced using specialized image analysis soft-
ware. The three-dimensional shape of the den-

Fig. 7 Cross section through an alloy 1100 ingot cast


by the Properzi (wheel-and-belt) method show- Fig. 8 Longitudinal section through 25 mm (1 in.) thick
ing columnar grains growing perpendicularly to the faces slab of alloy 1100 cast by the Hazelett (two-belt)
of the mold. Tucker’s reagent. Original magnification 1.5⳯ method. Tucker’s reagent. Actual size. Source: Ref 4

Fig. 6 Microstructure in cross and longitudinal sections


of DS Al 7050 alloy cooled at 0.45 ⬚C/s (0.8 ⬚F/
s) (BSE images). The actual bulk composition: 2.60% Cu,
2.37% Mg, 6.56% Zn, 0.03% Cr, 0.09% Fe, 0.05% Mn,
0.01% Ni, 0.06% Si, 0.018% Ti, 0.10% Zr, balance alu- Fig. 9 Cross section of a 150 mm (6 in.) diam ingot of alloy 6063 direct-chill semicontinuous cast. (a) Without grain
minum refiner. (b) With grain refiner. Tucker’s reagent. Actual size. Source: Ref 4
Name /bam_asmint_106420/6044_002d5/Mp_4 07/21/2004 02:27PM Plate # 0 pg 4 #
4 / Metallurgy and Microstructure

drites can be observed in the microshrinkage matic third-generation synchrotron radiation and with this technique are shown in Fig. 25. Two
cavities of the fractured specimen, as shown in high-resolution fast-readout detector systems grains nucleated through the detachment mech-
Fig. 24. have been used for in situ studies of dendritic anism are shown: one in the center of the figures
Recently, time-resolved direct-beam x-ray im- growth processes in aluminum-copper alloys and one on the right (marked with arrow). The
aging with intense, coherent, and monochro- (Ref 16). Two frames from a movie produced

Fig. 16 Silicon crystals with flakelike morphology in an


unmodified sample of alloy A365. Scanning
Fig. 10 Longitudinal section through a 75 mm (3 in.) Fig. 13 Outlining of grains using chemical etching.
electron micrograph after deep etching. Original magnifi-
diam alloy 1100 ingot, direct-chill cast without Etchant: modified Poulton reagent (60% HCl,
grain refiner. Center of section contains fan-shaped zones 30% HNO3 5% HF, 5% H2O) cation 1000⳯. Source: Ref 2
of feather crystals. Tucker’s reagent. Actual size. Source:
Ref 4

Fig. 14 Outlining of grains using thermal etching. Sam-


ples held for 1 h at 590 ⬚C (1095 ⬚F), then
quenched in cold water and dried

Fig. 17 Eutectic silicon crystals with fibrous morphol-


Fig. 11 Feather crystals in an alloy 3003 ingot cast by ogy in a modified sample of alloy A365. Etch-
the direct-chill semicontinuous process. ant: Keller’s reagent
Growth twins in the crystals. Polarized light. Barker’s re-
agent. Original magnification 50⳯. Source: Ref 4

Fig. 12 Macrostructure of an Al-12.7Si alloy showing Fig. 18 Silicon crystals with seaweedlike morphology
equiaxed grains and dendrites. Etchant: modi- Fig. 15 Eutectic silicon crystals with acicular mor- in a modified sample of alloy A365. Scanning
fied Poulton reagent (60% HCl, 30% HNO3 5% HF, 5% phology in an unmodified sample of alloy electron microscopy image after deep etching. Original
H2O). Original magnification 5⳯ A365. Etchant: Keller’s reagent magnification 1000⳯. Source: Ref 2
Name /bam_asmint_106420/6044_002d5/Mp_5 07/21/2004 02:27PM Plate # 0 pg 5 #
Solidification Structures of Aluminum Alloys / 5

−0.108
images are taken 34 s apart, and growth of the positions, the effect of increasing solute content SDAS = 14.15 ⋅ CMg ⋅ CSi−0.125 ⋅ tf0.302
new crystals is clearly evident. at a constant cooling rate is to decrease the den- −0.235
SDAS = 65.21 ⋅ CMg ⋅ CSi−0.104 ⋅ TL−0.22
The secondary dendrite arm spacing (SDAS) drite arm spacing (Fig. 28) (Ref 18).
depends on the local solidification time and cool- For Al-Cu-Si alloys the change in the copper
ing rate as discussed in the article “Fundamentals content of the alloy has greater effect on the Microsegregation
of Solidification” in this Volume. For Al-4.5Cu SDAS than silicon, when silicon content is
alloys, the experimental data (Ref 17) in Fig. 26 greater than 0.5%; otherwise the silicon has Most of the major alloying additions in alu-
can be fitted to Eq 7 in that article when the greater effect. The combined influence of local minum are less soluble in the solid phase than
coarsening constant is 10ⳮ6 m/s2. The influence solidification time, tf, cooling rate at the liquidus in the liquid phase (equilibrium partition ratio, k
of cooling rate on SDAS is illustrated in Fig. 27. temperature, ṪL, copper composition, CCu, and ⬍ 1). Moreover, for most solutes, aluminum ex-
The influence of solute content is less well silicon composition, CSi, on the SDAS can be hibits a relatively low terminal solid solubility;
defined (Ref 4). In general, up to eutectic com- calculated with (Ref 19): therefore, second-phase constituents are invari-
−0.333
ably present in cast structures. For these reasons,
SDAS = 34.9 ⋅ CCu ⋅ CSi−0.145 ⋅ tf0.208 the dendrites, which are the first portions of a
−0.1989
SDAS = 77.09 ⋅ CCu ⋅ CSi−0.078 ⋅ TL−0.252 cast structure to solidify, are low in solute con-
tent and are surrounded by interdendritic net-
For Al-Mg-Si alloys, changing the magne- works of one or more second-phase constituents.
sium content of the alloy has a greater effect on The size and distribution of the constituents de-
the SDAS than changing the silicon content, for pend on such factors as solute concentration,
silicon content greater than 0.5%. The combined dendrite arm spacing, and grain size. The solute
influence of local solidification time, cooling rate distribution characteristic of cast alloys may be
at the liquidus temperature, magnesium compo- described with reasonable accuracy by the Scheil
sition, CMg, and silicon composition, CSi, on the equation (see Eq 3 in the article “Fundamentals
SDAS can be calculated with (Ref 20): of Solidification” in this Volume). Examples of
elemental microsegregation are shown in Fig. 29
for the case of silicon microsegregation between
the ␣-aluminum dendrites of solid solution and
the eutectic in an Al-12.7Si cast sample, and in
Fig. 30 for the case of copper and magnesium
microsegregation in a 2124 direct-chill cast in-
got. Microsegregation can be decreased by a ho-
mogenization treatment.

Macrosegregation

The chemical composition of large, commer-


cial-size ingots can vary significantly from point
to point through the ingot thickness, which is
usually greater than 406 mm (16 in.) (see Fig.
31). This type of segregation, referred to as “ma-
crosegregation,” is only slightly affected by ho-
mogenization. In general, macrosegregation can
be reduced in direct-chill ingot casting by de-
creasing ingot thickness, lowering the casting
speed, and maximizing molten metal superheat
Fig. 20 Adsorption of impurity atoms on growth steps
(Ref 4). Macrosegregation is less of a problem
Fig. 19 Location of twin planes and re-entrant edge in of a silicon crystal causing twinning. Source:
a silicon crystal. Source: Ref 2 Ref 12 in shaped castings. The mechanisms of forma-

Fig. 21 Modified as-cast microstructures of aluminum alloys. (a) Al-10Si, Strontium-modified eutectic. Source: Ref 13. (b) Alloy A356.0, strontium-modified eutectic. Source: Ref
14. (c) Alloy A356.0, sodium-modified eutectic. Original magnification 100⳯. Source: Ref 14
Name /bam_asmint_106420/6044_002d5/Mp_6 07/21/2004 02:27PM Plate # 0 pg 6 #
6 / Metallurgy and Microstructure

Fig. 22 Alloy A356.0 near-quenched eutectic growth interface. (a) Secondary electron image. (b) Electron backscat-
tering diffraction map. See Fig. 2D5-22 on page XXX of the article “Selected Color Images” in this Volume.
Colors indicate the crystallographic orientation as represented in (c). Source: Ref 13

Fig. 25 Dendritic growth and grain multiplication in


Fig. 24 Scanning electron micrography image of alu- Al-20Cu alloy. Image (b) was taken 34 s after
minum dendrites in the fractured surface of a image (a). Arrow indicates grains nucleated by detachment.
tensile test bar of an A356.0 alloy Source: Ref 16

The size and shape of the pore is determined of gas rejected by the melt, pores form in the
Fig. 23 Typical dendrites in an A356 alloy in a com- by the amount of gas rejected by the melt. If this later stages of solidification, after dendrite co-
puter-processed image. Etchant: modified amount is large, spherical pores will form and herency. The concomitant development of
Poulton reagent (60% HCl, 30% HNO3 5% HF, 5% H2O) continue to grow through diffusion before den- shrinkage cavities nucleated by these pores re-
drite coherency is reached (Fig. 32). Their size sults in microshrinkage. The appearance of mi-
can be of the order of tenths of mm to mm. This croshrinkage is not spherical, but follows the
tion of macrosegregation are discussed in detail defect is called gas porosity. At lower amounts dendrite shape (Ref 22, 23) as seen in Fig. 33(a)
in Ref 21.

104
Defects

Macroporosity and Microshrinkage. Gas 103


evolution during solidification is responsible for
Dendrite arm spacing, µm

at least two casting defects: macroporosity and 102


microshrinkage (also called “microporosity” in
the literature). While both defects have a signifi-
cant influence on mechanical properties (ductil- 10
ity, fatigue), their mechanism of formation is
quite different. Macroporosity results when gas
is rejected from the liquid and is entrapped in the 1
solidifying metals as spherical gas defects of
millimeter size. Microshrinkage occurs when
10–1
liquid metal cannot reach interdendritic areas
during casting solidification. It is caused by a
combination of shrinkage and gas evolution, be- 10–2
ing the standard foundry defect for mushy-freez- 10–4 10–3 10–2 10–1 1 10 102 103 104 105 106 107
ing alloys, such as aluminum alloys, steel, su-
Local solidification time, s
peralloys, brass, bronze, and cast iron.
Microshrinkage voids are of the order of 10 to Fig. 26 Relation between secondary dendrite arm spacing and solidification time for Al-4.5Cu alloys. Data from 10
100 lm (Ref 21). investigators. Source: Ref 17
Name /bam_asmint_106420/6044_002d5/Mp_7 07/21/2004 02:27PM Plate # 0 pg 7 #
Solidification Structures of Aluminum Alloys / 7

and (b). If the melt is very well degassed, the gas solved in the melt. Unless the gas content of the dence demonstrates that a high level of
bubbles form at the very end of solidification liquid is reduced below the solid solubility of the impurities in the melt is associated with a high
between the eutectic grains (Fig. 33c). gas prior to solidification, the gas will precipitate microporosity.
The amount of microshrinkage seems to de- when the last liquid to freeze (usually the eutec- Surface Defects. A surface layer containing
pend directly on the amount of hydrogen dis- tic) solidifies. an undesirable concentration of alloying ele-
Modification of aluminum-silicon alloys with ments and associated coarse particles is often
sodium or strontium increases the volume frac- found in direct-chill semicontinuous cast alu-
tion of microshrinkage (Ref 24). This has been minum alloy ingots (Ref 4). A typical example
explained through a number of hypotheses such in an alloy 7075 ingot is shown in Fig. 34. The
as decreased liquid surface tension because of constituents segregated near the surface are
modification, change in the dendrite morphol- mainly Al-MgZn2, iron-containing phases, and
ogy, increased number of inclusions, and ob- MgSi2. In alloys with a large mushy zone, such
struction of the liquid flow. Experimental evi- as alloy 7075, exudations may occur on the sur-
face. A similar surface defect, associated with
dilute alloys, is the presence of bleed bands (Ref

Fig. 29 Secondary electron microscopy image of den-


drites and eutectic from an Al-12.7Si cast spec-
imen. Silicon microsegregation between the dendrites of
solid solution and eutectic revealed by electron probe mi-
croanalysis

Fig. 31 Variation in copper concentration across a 600


mm (24 in.) thick direct-chill semicontinuous
cast ingot of 2124 alloy. Source: Ref 4

Fig. 27 Microstructures of A356 alloy solidified at dif-


ferent cooling rates. (a) Cast in metallic mold
(high cooling rate), fine dendrites and network of interden-
dritic eutectic form. (b) Cast in green sand mold (low cool-
ing rate), coarse dendrites and discontinuous network of
interdendritic eutectic result. Etchant: Keller’s reagent

Fig. 30 Microstructure at a midthickness location. Di-


rect-chill semicontinuous cast 610 ⳯ 1372 Fig. 32 Pores formed by gas evolution in the casting.
mm (24 ⳯ 54 in.) 2124 alloy ingot. Etchant: (a) 0.5% HF. (a) Optical micrograph of a pore in aluminum-
Fig. 28 Effect of copper content on secondary arm (b) Copper and magnesium microsegregation (revealed by silicon alloy. Etchant: 0.5% HF. (b) Scanning electron mi-
spacing in eight aluminum alloys, plotted for electron probe microanalysis) across the dendrites. Source: croscopy image of pore in unmodified aluminum-silicon
five cooling rates. Source: Ref 18 Ref 4 alloy
Name /bam_asmint_106420/6044_002d5/Mp_8 07/21/2004 02:27PM Plate # 0 pg 8 #
8 / Metallurgy and Microstructure

Fig. 33 Microshrinkage in aluminum alloys. (a) Optical micrograph of interdendritic microshrinkage in aluminum-silicon alloy. Etchant: 0.5% HF. (b) Scanning electron microscopy
image of interdendritic microshrinkage in aluminum-silicon alloy. (c) Scanning electron microscopy image of microshrinkage between the eutectic grains in aluminum-
silicon alloy

25) (see Fig. 35). Both defects are caused by may extend 20 mm (0.75 in.) below the chilled such as that observed in Fig. 37(b) are seen. They
reheating the solidified shell of the casting when surface (Ref 26). The contrast between the exhibit a torn surface indicating that they have
it separates from the mold wall, briefly ceasing coarse structure of the surface layer and the fine probably been formed by deformation and neck-
heat removal (Ref 4). subsurface structure of an alloy 2024 ingot is ing of a solid bridge extending across the grain
The surfaces of aluminum alloy ingots are of- shown in Fig. 36. boundary (Ref 27).
ten “scalped” to remove segregated or nonuni- Hot tearing is the result of tensile stresses Columnar structures are particularly prone to
form surface layers, which vary in thickness and building up in the mushy zone at the end of so- hot tearing, which may occur at the intersection
lidification. The mass deficit in the last volume of two columnar fronts, because of the mass def-
to solidify produces voids in the interdendritic icit in the last region to solidify, and because of
liquid as the dendrites are pulled apart when fluid a buildup of internal tension stresses during final
flow cannot compensate for shrinkage. This de- cooling of ingots. An example of center cracking
fect is most common in dilute alloys. in an alloy 1100 ingot is shown in Fig. 38. Center
The hot crack surface is very smooth, with cracks (at arrow) resulted from excessively steep
small undulations corresponding to secondary temperature gradients.
dendrite arms (Fig. 37). This seems to indicate Inclusions and Oxide Films. According to
that hot cracking occurred while the grain recent research (Ref 28), liquid metals cannot
boundary was covered by a thin liquid film. At
many locations, spikes such as those shown in
Fig. 37(a) can be observed on both sides of the
hot cracked surface. They seem to have formed
by elongation of a liquid region with simulta-
neous oxide formation at the surface, but without
appreciable deformation. In a few cases, spikes

Fig. 34 Section through an alloy 7075 ingot (edge at


right), direct-chill semicontinuous cast. Etch-
ant: dilute Keller’s reagent. Original magnification 250⳯.
Source: Ref 4

Fig. 36 Section through a 305 mm (12 in.) diam alloy


Fig. 35 Bleed bands normal to the casting direction on 2024 ingot, direct-chill semicontinuous cast. Fig. 37 Scanning electron microscopy observation of
the surface of a direct-chill semicontinuous Coarse dendrites and copper and magnesium constituents Al-3Cu. (a) Spike probably formed by the last
cast ingot of alloy 3003. Unpolished. One-fourth actual agglomeration in the surface layer. As-polished. Original solidification of interdendritic liquid. (b) Deformed spike
size. Source: Ref 25 magnification 25⳯. Source: Ref 4 probably formed by necking of a solid bridge
Name /bam_asmint_106420/6044_002d5/Mp_9 07/21/2004 02:27PM Plate # 0 pg 9 #
Solidification Structures of Aluminum Alloys / 9

Fig. 38 Cross section through a 75 mm (3 in.) diam Fig. 41 Scanning electron microscopy image of oxide
inclusions in aluminum cast samples (fractured
alloy 1100 ingot, direct-chill cast. Center surface)
cracks at arrow. Etchant: Tucker’s reagent. Actual size.
Source: Ref 4

Ohnaka and D.M. Stefanescu, Ed., 1996, p


55–63
12. S.-Z. Lu and A. Hellawell, Met. Trans. A,
Vol 18A, Oct 1987, p 1721–1733
13. K. Nogita and A.K. Dahle, Scr. Mater., Vol
48, 2003, p 307–313
14. J.E. Gruzleski and B.M. Closset, The Treat-
ment of Liquid Aluminum-Silicon Alloys,
AFS, 1990
Fig. 40 Coarse primary crystal of CrAl7 in an alloy
15. G. Heiberg, K. Nogita, A.K. Dahle, and L.
7075 ingot. As-polished. Original magnifica-
tion 100⳯. Source: Ref 4 Arnberg, Acta Mater., Vol 50, 2002, p 2537–
Fig. 39 Surface turbulence, acting to fold in a bifilm
2546
and bubbles. Source: Ref 28
16. R. Mathiesen, L. Arnberg, K. Ramsoskar, T.
Weitkamp, C. Rau, and A. Snigirev, Metall.
nucleate such defects as pores and cracks by het- Other inclusions may be present in the solid- Mater. Trans. B, Vol 33B, 2002, p 613–623
erogeneous nucleation. The melt surface is usu- ifying aluminum casting (Fig. 40). Hard inclu- 17. M.C. Flemings, T.Z. Kattamis, and B.P.
ally covered with a film, usually an oxide film. sions may perforate the oxide films and initiate Bardes, AFS Trans., Vol 99, 1991, p 501
The entrainment of this surface film during pour- fracture (Fig. 41). 18. J.A. Horwath and L.F. Mondolfo, Acta Me-
ing leads to many defects on which pores and tall., Vol 10, 1962, p 1037–1042
cracks can grow. The defects include bubbles REFERENCES 19. V. Ronto and A. Roosz, Int. J. Cast Met.
and their associated bubble trails of various Res., Vol 13, 2001, p 337–342
kinds, and folded films (bifilms) that act as 1. I.J. Polmear, Light Metals, E. Arnold Pub- 20. V. Ronto and A. Roosz, Int. J. Cast Met.
cracks and as substrates for the precipitation of lishers, London, 1981 Res., Vol 14, 2001, p 131–135
many phases. Most failure mechanisms of cast 2. L. Bäckerud, G.C. Chai, and J. Tamminen, 21. D.M. Stefanescu, Science and Engineering
products appear to be associated with bifilms, Solidification Characteristics of Aluminum of Casting Solidification, Kluwer Academic,
particularly those introduced during mold filling. Alloys, Vol 2, AFS/Skanaluminum, 1990 2000
This seems to be particularly true of hot tearing, 3. F. Xie, X. Yan, L. Ding, F.Z., S. Chen, M.G. 22. R. Fuoco, H. Goldenstein, and J.E. Gru-
which appears to be eliminated in even the most Chu, and Y.A. Chang, Mater. Sci. Eng. A, zleski, AFS Trans., Vol 101, 1994, p 297
susceptible alloys simply by reduced surface tur- Vol A00, 2003, p 1_/10 23. T.S. Piwonka, Proc. Merton C. Flemings
bulence during mold filling. 4. D.A. Granger, Solidification Structures of Symposium on Solidification and Materials
The entrainment mechanism is a simple fold- Aluminum Alloy Ingots, Metallography and Processing, R. Abbaschian, H. Brody, and
ing-in action, thus folding the surface oxide dry Microstructures, Vol 9, ASM Handbook, A. Mortensen, Ed., TMS, 2000, p 363
side to dry side, and entraining a thin layer of ASM International, 1985, p 629–636 24. D. Argo and J. Gruzleski, AFS Trans., Vol
adsorbed air between the films (see Fig. 39) 5. R.E. Spear, R.T. Craig, and C.R. Howle, J. 96, 1988, p 65
forming a bifilm. Since there is practically no Met., Vol 23 (No. 10), 1971, p 42–45 25. D.L.W. Collins, Metallurgia, Vol 76, Oct
bonding across the air layer bounded by these 6. A. Cibula, J. Inst. Met., Vol 80, 1951, p 1– 1967, p 137–144
dry interfaces, the defect acts as a crack in the 16 26. G. Siebel, D. Altenpohl, and M.H. Cohen,
liquid (Ref 28). The bifilms can be extremely 7. L. Bäckerud, Jernkontorets Ann., Vol 155, Z. Metallkde., Vol 44, 1953, p 173–183
thin: the oxides may be only 20 nm thick. If the 1971, p 422–424 27. M. Rappaz, J.M. Drezet, P.D. Grasso, and
air layer has a similar thickness the total thick- 8. L. Bäckerund, Light Met. Age, Oct 1983, p A. Jacot, Modeling of Casting, Welding and
ness may be only 60 nm, although total thickness 6–12 Advanced Solidification Processes, D.M.
in the range 1 to 10 lm is common. However, 9. L.F. Mondolfo, Light Metals, Metallurgical Stefanescu, J. Warren, M. Jolly, and M.
the area of the defect can be large. In aluminum Society of AIME, 1972, p 405–426 Krane, Ed., TMS, 2003, p 53–60
castings 10 mm square is common, but 10 to 10. D.A. Granger and J. Liu, J. Met., Vol 35 28. J. Campbell, Modeling of Casting, Welding
1000 lm is probably more usual. Their volume (No. 6), 1983, p 54–59 and Advanced Solidification Processes,
density (number per mL) can range from 1 to 11. C. Degand, D.M. Stefanescu, and G. Laslaz, D.M. Stefanescu, J. Warren, M. Jolly, and
1000 or more. Solidification Science and Processing, I. M. Krane, Ed., TMS, 2003, p 209–218

View publication stats

You might also like