Vous êtes sur la page 1sur 10

Materials Science and Engineering A 480 (2008) 496–505

Aging behaviour of Al–Cu–Mg alloy–SiC composites


Sharmilee Pal a , R. Mitra a,∗ , V.V. Bhanuprasad b
aDepartment of Metallurgical and Materials Engineering, Indian Institute of Technology, Kharagpur 721302, W.B., India
b Ceramics and Composites Group, Defence Metallurgical Research Laboratory, P.O. Kanchanbagh, Hyderabad 500058, A.P., India
Received 5 April 2007; received in revised form 11 July 2007; accepted 26 July 2007

Abstract
The age-hardening kinetics of powder metallurgy processed Al–Cu–Mg alloy and composites with 5, 15 or 25 vol.% SiC reinforcements,
subjected to solution treatment at 495 ◦ C for 0.5 h or at 504 ◦ C for 4 h followed by aging at 191 ◦ C, have been studied. The Al–SiC interfaces
in composites show undissolved, coarse intermetallic precipitates rich in Cu, Fe, and Mg, with its extent varying with processing conditions.
Examination of aging kinetics indicates that the peak-age hardness values are higher, and the time taken for peak aging is an hour longer on
solutionizing at 504 ◦ C for 4 h, due to greater solute dissolution. Contrary to the accepted view, the composites have taken longer time to peak-age
than the alloy, probably due to lower vacancy concentration, large-scale interfacial segregation of alloying elements, and inadequate density of
dislocations in matrix. The composite with 5 vol.% SiC with the lowest inter-particle spacing has shown the highest hardness.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Discontinuously reinforced aluminum composites; Aging; Kinetics; Interfaces; Hardness

1. Introduction have reported slower kinetics of aging in the powder metallurgy


processed, liquid phase sintered Al–SiC composites with coarse
Discontinuously reinforced aluminum (DRA) matrix com- and widely spaced SiC particles, compared to that of similarly
posites have received significant attention in recent years. The processed unreinforced Al alloy, and attributed its cause to rel-
high specific strength and stiffness make the DRA composites atively low dislocation density in the matrix. In general, it is
attractive as candidate materials for aerospace and automo- accepted that the aging behaviour depends on the reinforcement
tive applications. It has been reported that the dislocations are and its volume fraction, alloy composition, heat treatment, and
punched out at the reinforcement–matrix interfaces to relax other processing parameters.
the stress generated owing to the mismatch in coefficients of The present study is aimed at examination of the role of SiC
thermal expansion (CTE) [1,2] of Al (24 × 10−6 K−1 ) and SiC reinforcement particles and interfaces on the aging kinetics of
(4 × 10−6 K−1 ). It has also been reported that the age-hardenable the matrix in a powder metallurgy (P/M) processed Al–Cu–Mg
Al alloy–SiC composites demonstrate accelerated kinetics of alloy–SiC composite, for two different conditions of solution
aging [3–13], when compared to that of the unreinforced alloy treatment, leading to different degrees of dissolution of matrix
due to the much greater dislocation density in the former. The precipitates.
dislocations provide sites for heterogeneous nucleation of the S
(CuMgAl2 ) precipitates [7]. Suresh et al. [6] have concluded that 2. Experimental procedure
the volume fraction of SiC reinforcements beyond 6% has no sig-
nificant effect on the aging kinetics of Al–3.5Cu alloy matrix in The experimental procedure involves processing and heat
the cast composites. However, Lin et al. [10] have shown a mono- treatment of the composites, and characterization of microstruc-
tonic decrease in the peak-aging time for the 2024 Al alloy matrix ture and interfaces.
with increasing volume fraction of SiC reinforcement particles.
Contrary to the results of most studies, Janowski and Pletka [14]
2.1. Processing

∗ Corresponding author. Tel.: +91 3222 283292; fax: +91 3222 282280. The Al–Cu–Mg alloy–SiC composites for the present investi-
E-mail address: rahul@metal.iitkgp.ernet.in (R. Mitra). gation were obtained from the Ceramics and Composites group,

0921-5093/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.07.072
S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505 497

Defence Metallurgical Research Laboratory, Hyderabad, India. electron microscope (TEM), by mechanical thinning, followed
Through repeated bulk EDX analyses and optical emission spec- by jet polishing using an electrolyte comprising 33 vol.% HNO3
troscopy, the Al–Cu–Mg alloy matrix has been found to contain and 67% methanol in temperature range of −20 to −30 ◦ C. Sub-
5.4 Cu, 1.5 Mg, 0.3 Si, 0.02 Mn, 0.4 Fe, 0.07 Zn, 0.2 Ti, 0.2 Cr, sequently, TEM observations in bright and dark field imaging
and balance Al (all in wt.%). modes accompanied by energy dispersive X-ray analyses were
The composites were processed by vacuum hot pressing carried out on a JEOL 2100 HRTEM operating at an acceleration
of the intimate mixtures of inert gas atomized powders of voltage of 200 kV.
Al–Cu–Mg alloy (Metal Powder Company Limited, Madurai) In order to study the kinetics of aging, room temperature
and commercial ␣-SiC particles (Carborundum, India), followed microhardness measurements were performed on the matrix of
by extrusion at the ratio of 25:1 after soaking at 450 ◦ C for the composites in as-extruded and differently aged conditions
0.5 h. The process of extrusion helps in a more uniform dis- using a Vickers diamond indenter at loads of 10, 25, 50 or 100 gf
tribution of SiC particles. The steps involved in processing have on a computerized LECO microhardness tester. At least 10 mea-
been described in detail elsewhere [15]. The average size of SiC surements were recorded at each load for every specimen and
particles was 5 ␮m in the composite reinforced with 5 vol.% averaged. Vickers hardness of the differently aged specimens
SiC, and 50 ␮m in case of those with 15 and 25 vol.% SiC. was also measured using the 10 kg load.
The as-extruded Al–Cu–Mg alloy–SiC composites were solu- The Al alloy–15 vol.% SiC composite sample, solution
tion treated either at 495 ◦ C for 30 min, or at 504 ◦ C for 4 h and treated at 500 ◦ C (500 ◦ C is almost the average of 495 and
quenched in ice-cold water, following the procedures for alloys 504 ◦ C) for 0.5 h was electropolished in an electrolytic bath
of similar compositions mentioned elsewhere [5,16]. The tem- containing phosphoric acid, ethanol, and water [19] and sub-
perature variation during each solution treatment was ±2 ◦ C. sequently etched to reveal the dislocations using an etchant
The conditions of solution treatment represent two extremes of containing a mixture of HCl, HNO3 and HF [20]. The dislo-
the expected degree of dissolution of solute in the matrix. For cation density has been measured using image analysis of etch
solution treatment, the specimens were placed inside the fur- pits, and compared with that obtained through theoretical calcu-
nace, after the required temperature was reached, and time was lations. A large area of the specimen could be observed without
measured only after the temperature stabilized. The cooling rate the artifacts, generally introduced by deformation during TEM
is expected in the range of 475–480 ◦ C/s (K/s) during quench- sample preparation. The method of etching is usually applicable,
ing, assuming a momentary fall of 15–20 ◦ C in the time span when dislocation density is relatively low.
between exit from the furnace and dropping into the ice-cold
water. Some of the solution-treated samples were refrigerated 3. Results and discussion
(≈−15 ◦ C) inside the freezer for further characterization, while
others were subsequently subjected to artificial aging at 191 ◦ C The microstructures of the Al–Cu–Mg alloy and its compos-
for various lengths of time. For the purpose of aging, the sam- ites with 5, 15, and 25 vol.% SiC reinforcements, in as-extruded,
ples were inserted into the furnace after the aging temperature solution treated, or aged conditions have been studied with
was reached, and only a momentary drop of roughly 20 ◦ C was emphasis on the nature and location of the precipitates, and
noticed during insertion. The Al–Cu–Mg alloy matrix will be discussed in this section.
referred to as the “Al alloy” henceforth.
3.1. Microstructural examination
2.2. Characterization
3.1.1. As-extruded condition
The microstructures and particle–matrix interfaces of the Fig. 1 is a typical SEM (SE) image of the as-extruded Al
as-extruded and heat-treated samples were studied using sec- alloy–25 vol.% SiC composite, showing a uniform distribution
ondary (SE) and back-scattered electron (BSE) imaging modes of SiC particles, and is representative of the other composites
on a JEOL JSM-5800 scanning electron microscope (SEM). as well. SEM examination has indicated that the 15 or 25 vol.%
The chemical compositions of the different phases were inves- SiC reinforced composites also show a small fraction of particles
tigated through spot and line-scan analyses on Link AN 10/55s with size in the range of 5–10 ␮m, much less than the average
Energy Dispersive X-Ray (EDX) detector, attached to the SEM. of 50 ␮m. The smaller size particles owe their origin to cracking
The particle sizes were measured on an optical microscope, during extrusion. The typical XRD patterns of the as-extruded
and from SEM micrographs using an image analysis software composite with 5, 15 or 25 vol.% SiC particles (Fig. 2) show the
along longitudinal and transverse directions, following the pro- peaks of precipitate phases, CuAl2 and CuMgAl2 (S ), besides
cedure recommended in the ASTM E112-96 standard [17]. The those of Al and SiC. Fig. 3(a) and (b) shows the typical SEM
interface area was measured from the sum of perimeters of the (BSE) images depicting the microstructure of as-extruded com-
SiC particles, or the total interface length per unit area, using posites with 5 and 25 vol.% SiC, respectively. The precipitates
the intercept method proposed by Underwood [18]. To iden- could be observed distinctly in all the composites in the SEM
tify the phases formed during the aging heat treatments, X-ray (BSE) images [Figs. 3(a) and (b)], as those appear bright due
diffraction (XRD) studies were conducted on a Philips PW1840 to the higher atomic number of Cu or Fe, compared to those of
diffractometer using Cu K␣ radiation. Electron transparent spec- Al, Si or C. Fig. 3(a) and (b) also shows that a large fraction of
imens were prepared, for observations on the transmission the particle–matrix interfaces has precipitates. EDX analyses of
498 S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505

Fig. 1. SEM SE image of Al alloy–25 vol.% SiC composite, showing a uniform


distribution of particles.

the spots on interfacial precipitates have shown the peaks of Cu,


Mg and Fe, as is obvious from the typical spectrum depicted in
Fig. 3(c). Of course, there are several particle–matrix interfaces
without any precipitate.

3.1.2. Solution-treated condition


Studies on the composites with the matrix solution treated
at 495 ◦ C have shown undissolved precipitates at some of the
particle–matrix interfaces, as is evident from the SEM (BSE)
image in Fig. 4(a). Quantitative analysis considering about 50
SiC particles has shown that almost 50% of the SiC parti-
cles exhibits undissolved precipitates at their interfaces with
the matrix. EDX spot analyses, with spectra resembling that
in Fig. 3(c), have confirmed the enrichment of solute, particu-
larly Mg, Cu or Fe at many of the Al–SiC interfaces shown in
the SEM (BSE) image of Fig. 4(a). Magnified view of the XRD
pattern from the sample solution treated at 495 ◦ C is shown in
Fig. 4(b), and depicts very tiny peaks of undissolved CuAl2 or
CuMgAl2 precipitates. Thus the presence of precipitates at the

Fig. 3. As-extruded Al alloy–SiC composites: SEM (BSE) images with: (a)


5 vol.%, and (b) 25 vol.% reinforcements, showing precipitates present in
the matrix and at interfaces, and (c) EDX spectrum of a precipitate at the
particle–matrix interface, showing the peaks of Al, Cu, and Fe.

Fig. 2. XRD pattern of as-extruded Al alloy–25 vol.% SiC composite, showing


the peaks of CuAl2 and CuMgAl2 precipitates.
S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505 499

Fig. 5. SEM (BSE) image showing the microstructure of the Al alloy–25 vol.%
SiC composite solution treated at 504 ◦ C for 4 h.

dissolution of precipitates in the composites solution treated at


504 ◦ C for 4 h is closer to its completion than in those solution
treated at 495 ◦ C for 0.5 h. It is imperative that both the time and
temperature of solution treatment have an effect on the degree of
dissolution. The root mean square diffusion distance of a given
solute atom under non-steady state condition is proportional to
Dt, where D is the diffusion coefficient and t is the time [21]. The
term, D is related to the absolute temperature by an Arrhenius
type relation. In the present study, the difference in temperature
is only 9 ◦ C, but time of solution treatment at the 504 ◦ C is eight
times that at 495 ◦ C. The effects of temperature and time could be
distinguished from one another, only through further systematic
studies on solution treatment, keeping one of those unchanged
Fig. 4. Al alloy–25 vol.% SiC composite solution treated at 495 ◦ C for 0.5 h: (a) and the other varying.
SEM (BSE) image of the microstructure showing precipitates at the interface Negligible solubility of Fe in Al does not allow dissolution
and in the matrix; and (b) XRD pattern showing peaks of precipitates.
at either of the solution treatment temperatures. In addition, the
atoms of Mg have a strong tendency to segregate at interfaces, as
interfaces, and small peaks of CuAl2 or CuMgAl2 , appearing in confirmed through EDX spot analyses. The solute atoms tend to
the XRD patterns is suggestive of incomplete dissolution dur- segregate at the interfaces in the composites in solution-treated
ing solution treatment. EDX spot analysis from matrix of the condition, since the free energy is reduced by Gibbs adsorp-
composite solution treated at 495 ◦ C for 0.5 h has also shown tion. In certain other studies, the surfaces of SiC particles have
Cu concentration as low as 2–3 wt.%, which is much less than also provided preferred sites for the nucleation of intermediate
that in the bulk. precipitates [22,23]. At high cooling rates used in this study,
On the other hand, the concentration of Cu in the alloy matrix the possibility of formation of intermediate precipitates is also
on solution treatment at 504 ◦ C has been found to be almost unlikely [24].
similar to the bulk composition. Fig. 5 is an SEM (BSE) image The density of dislocations appearing as etch-pits in the
of the composite with 15 vol.% SiC particles solution treated at microstructure of the Al alloy–15 vol.% SiC composite (Fig. 6)
504 ◦ C for 4 h, showing undissolved precipitates at a very limited with matrix in solution-treated condition has been found to be
number of particle–matrix interfaces. EDX spot analyses have 7 × 1011 m/m3 .
confirmed the enrichment of Mg and Fe in the coarse precipitates
at the SiC–matrix interfaces in the alloy solutionized at 504 ◦ C 3.1.3. Aged condition
for 4 h. Quantitative analysis considering about 50 SiC particles Fig. 7(a–c) shows the SEM (BSE) images of the Al alloy–5,
in the SEM (BSE) images, has shown precipitates at about 10% 15, and 25 vol.% SiC composites, over-aged after solution treat-
of the interfaces. Thus a comparison of the SEM (BSE) images ment at 495 ◦ C for 0.5 h. A large number of the SiC–matrix
in Figs. 4(a) and 5 shows that the undissolved precipitates occur interfaces show precipitates enriched in Cu, Mg or Fe. Since the
less frequently in the matrix and at interfaces, compared to that in Al alloy–SiC interfaces also have a higher density of dislocations
samples solution treated at 495 ◦ C for 0.5 h, confirming a much compared to that in the matrix, in addition to a higher degree of
larger extent of solute dissolution in case of the former. The interfacial solute enrichment, the nucleation and growth of pre-
500 S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505

Fig. 6. SEM (SE) image of the Al alloy–15 vol.% SiC composite, showing etch-
pits for dislocations present.

cipitates at and near interfaces is usually preferred during aging.


The XRD patterns from the underaged or overaged Al-alloy–SiC
composite have shown the peaks of CuAl2 and CuMgAl2 phases
as expected. The ternary Al–Cu–Mg phase diagram [25] shows
equilibrium between ␣-Al, ␪ (CuAl2 ), and S (CuMgAl2 ) phases.
Fig. 8(a) and (b) is the typical bright and dark field TEM
images of the Al alloy in under-aged condition, showing rod-
shaped precipitates of the CuMgAl2 phase. The selected area
diffraction (SAD) pattern from the area with the precipitates, is
shown as inset in the dark field TEM image of Fig. 8(b). The
diffraction spots are streaked perpendicular to the rod-shaped
precipitates. On examination of the bright and dark field TEM
images with reference to the diffraction pattern, it is obvious
that the axis of the CuMgAl2 rods are aligned in the orthogonal
1 0 0 direction of the matrix. The 1 0 0 orientation of the
rod-shaped CuMgAl2 precipitates in the TEM images [Fig. 8(a)
and (b)] agree with the observations previously reported in the
literature [26–29]. Fig. 8(c) is a typical diffraction pattern from
a larger area confirming the presence of CuMgAl2 as well as
CuAl2 precipitates in under-aged samples.

3.2. Theoretical analysis of interface area and


inter-particle spacing

The integrity of SiC–matrix interfaces and inter-particle spac-


ing strongly affect the dislocation density, which is known to
affect the aging behaviour in turn, as has been mentioned in
Section 1. Assuming the particles to be equiaxed or sphrerical, Fig. 7. SEM (BSE) images showing interfacial segregation: (a) Al alloy–5 vol.%
the net particle–matrix interface area per unit volume, AI (ignor- SiC composite after solution treatment and artificial aging for 16 h; (b) Al
ing the intermetallic precipitates in the alloy matrix) has been alloy–15 vol.% SiC composite after solution treatment and artificial aging for
48 h; and (c) Al alloy–25 vol.% SiC composite after solution treatment and
calculated using the relation [30]: artificial aging for 16 h.
6Vp
AI = (1)
d a function of volume fraction and particle size. Assuming the
where Vp is the volume fraction of particles and d is the particle particles to be equiaxed, the inter-particle spacing is calculated
size. Since the SiC particles are irregularly shaped, the calcu- using the following relationship [31]:
lated area of interfaces per unit volume is only approximate.
The average spacing between the dispersed SiC particles is also λ = 0.77d(Vp )−0.5 (2)
S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505 501

Fig. 8. TEM micrographs of underaged alloy: (a) bright and (b) dark field images showing lath-shaped precipitates of CuMgAl2 with the SAD pattern shown as the
inset; and (c) SAD pattern from a larger area showing rings representing polycrystalline Al, CuAl2 , and CuMgAl2 phases. The dark field image has been recorded
using the reflection, g = [2 0 0].

Table 1
The measured and calculated values of the particle-size, particle–matrix interface area, inter-particle spacing, and calculated value of the dislocation density in the
matrices of as-extruded Al alloy–SiC composites
Volume fraction of SiC (%) Particle size (␮m) Interface area (m2 /m3 ) Inter-particle spacing (␮m) Dislocation density, B (m/m3 )

Calculated Measured Calculated Measured

5 05 6.0 × 104 1.1 × 103 17.2 14.2 3.7 × 1011


15 50 1.8 × 104 4 × 104 99.4 95.7 1.2 × 1011
25 50 3.0 × 104 7 × 104 77.0 75.5 2.3 × 1011
502 S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505

Fig. 9. Plots showing the variation of matrix microhardness with time of aging for indentation loads of 10, 25, 50, and 100 gf in case of the (a) Al alloy, and composites
with (b) 5 vol.%, (c) 15 vol.%, and (d) 25 vol.% SiC particles. The solution treatment was carried out at 495 ◦ C for 0.5 h.

Table 1 shows the calculated and experimentally measured densities in the Al alloy–SiC composites, shown in the last col-
values of interface area per unit volume and the average inter- umn of Table 1, have the same order of magnitude in spite of the
particle spacing, obtained for composites with the different SiC different sizes and volume fractions of reinforcement particles,
particle sizes. The values of experimentally measured interface and increase with decreasing inter-particle spacing. The order
area are similar to those calculated from Eq. (1), except in case of magnitude of the theoretically evaluated dislocation densities
of the composites reinforced with 5 vol.% SiC particles with agrees well with that of experimental measurement, mentioned
5 ␮m average size. It may be noted that Eq. (1) was derived, in Section 3.1.2.
considering spherical particles, while the SiC particles are irreg-
ular in shape. Smaller the particle size, larger is the surface 3.3. Isothermal ageing kinetics
area to volume ratio. As a result, the deviation of experimen-
tally determined value of surface area from that evaluated using Fig. 9(a–d) shows the plots depicting the variation of matrix
Eq. (1) is maximum for the smallest particle size of 5 ␮m. On microhardness with time of aging obtained using indentation
the other hand, the experimentally determined values of inter- loads ranging between 10 and 100 gf, for the composites hav-
particle spacing have the same order of magnitude and are very ing 5, 15, and 25 vol.% SiC particles, respectively, after solution
close to those obtained through calculations using the relation- treatment at 495 ◦ C for 0.5 h. It is clear from the microhard-
ship (2). The composite with 5 vol.% SiC particles having the ness data, that the time taken for peak aging of the Al alloy
average size of 5 ␮m has shown the smallest inter-particle spac- and all the composites solution treated at 495 ◦ C are 3 and 5 h,
ing, which is close to the 1/5th of that between the particles in respectively. Fig. 10(a) and (b) shows the plots depicting the
the other two composites. aging kinetics (based on microhardness measurements at load
The density of dislocations with Burgers vector, b, generated of 100 gf) of the matrix in Al alloy and its composites with 5,
as a result of quenching from solution treatment temperature 15, or 25 vol.% SiC particles, previously subjected to solution
(500 ◦ C) to 0 ◦ C (ice-cold water) has been calculated using the treatment at 495 ◦ C for 0.5 h and 504 ◦ C for 4 h, respectively.
following Eq. [2]: The peak-aging condition could be reached in the monolithic
alloy and the composites solution treated at 504 ◦ C for 4 h after
(BVp ε) 4 and 6 h of aging, respectively. There is a very large scatter in
ρ= (3)
bt(1 − Vp ) the microhardness of the matrix, which could be due to contact
of the indenter with SiC particles in the sub-surface locations,
where B is a constant having value between 4 and 12, ε is the and a locally inhomogeneous distribution of reinforcements and
strain due to mismatch in CTE = (T) (CTE), and t is the small- precipitates. Different loads have been used to determine the
est dimension of the reinforcement. The calculated dislocation microhardness for better statistics and reliability, and the time
S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505 503

Fig. 11. Plots showing the variation of hardness of the Al alloy and its com-
posites, with duration of aging after solution treatment at 504 ◦ C for 4 h. The
hardness tests were carried out using a load of 10 kgf.

reduced kinetics of aging observed in the Al alloy–SiC compos-


ites is in contrast to what has been reported for most of the times
in the past [3–13].
The slower kinetics of aging in the Al alloy–SiC composites
compared to that in the alloy, could be due to one or more of the
following mechanisms:

(I) The presence of SiC particles leads to a reduction in the


volume fraction of GP zones, which follows from the lower
vacancy concentration in the composite, compared to that
in the unreinforced alloy [4,11,14,32]. The vacancy con-
centration is lower in the composites due to the large area
of particle–matrix interfaces, acting as vacancy sinks. It is
Fig. 10. Plots showing variation of matrix microhardness with time of aging, intuitive that the lower vacancy concentration in the matrix
after solution treatments at: (a) 495 ◦ C for 0.5 h, and (b) 504 ◦ C for 4 h, for the Al of the composite compared to that in the Al alloy is respon-
alloy and its composites with 5, 15, 25 vol.% SiC particles. The microhardness
sible for retarding the formation of GP zones. It has been
tests were carried out using a load of 100 gf.
argued in a previous study [11], that the initial retardation
of aging kinetics due to lower density of GP zones may
for reaching peak hardness has been found to be the same for affect the remaining steps of the aging sequence [14]. On
all. A smaller load gives a more localized microhardness with- the other hand, a higher concentration of vacancies in the
out interference from the neighbouring or sub-surface regions. Al alloy leads to a greater density of GP zones, enhancing
On the other hand, a larger load would cause a larger indenta- the aging kinetics in the initial stages. In a related study
tion covering a wider area. The variation of Vickers hardness of on the aging kinetics of the 2024 Al–9 or 14 vol.% Al2 O3
the Al alloy and the composites, solution treated at 504 ◦ C for fiber composites, Cen and Chao [32] have shown negligi-
4 h with time of aging is shown in Fig. 11. The aging kinetics ble effect of the reinforcement phase on the aging kinetics,
and time for peak aging determined through measurement of the and explained it on the basis of reduced vacancy concen-
microhardness [Fig. 10(b)] and Vickers hardness (Fig. 11) agree tration, and lower GP zone density restricting the S phase
well. formation as well.
It is obvious that the aging kinetics of the Al alloy is faster (II) Excessive segregation of alloying elements at the Al–SiC
than that of the composites for both the temperatures of solution interfaces and presence of undissolved, coarse precipi-
treatment [Fig. 9(a–d), 10(a and b), and 11]. Fig. 10(a) shows tates causing depletion of solute inside the matrix, found
that for solution treatment at 495 ◦ C for 0.5 h, the rate of increase particularly on solutionizing at 495 ◦ C for 0.5 h, leads to
in hardness with time of aging of the alloy is the highest, fol- lower supersaturation of solute. A lower degree of super-
lowed by that of composite with 5 vol.% SiC particles, while saturation would reduce the chemical driving force for
the composites with 15 and 25 vol.% SiC particles show a more precipitation in the composites. Previously, Hunt et al. [33]
gradual increase in hardness. The rise to peak hardness in less have reported retardation of S precipitation in matrix of
time is suggestive of the faster kinetics in the initial stages. The the composite, attributing its cause to higher intermetallic
504 S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505

content in the composite compared to that in the alloy, and pared to that at 504 ◦ C for 4 h is probably because the difference
incomplete solutionizing, which makes solute unavailable in the Cu concentration of the matrix between its solution treated
for S formation. and equilibrium condition is less in case of the former. In addi-
(III) The difference in kinetics of aging is caused by S pre- tion, the peak hardness of the alloy and composites on aging
cipitation at dislocations, density of which plays a critical after solution treatment at 504 ◦ C for 4 h is more than those of the
role. The dislocation density in the composites may be samples aged after solutionizing at 495 ◦ C for 0.5 h [Fig. 10(a)
insufficient for enhancement of kinetics of aging: and (b)], which again can be explained on the basis of the lower
(a) The experimentally determined and calculated dislo- volume fraction of precipitates in case of the latter caused by
cation density of the order of 1011 m/m3 (Table 1) incomplete dissolution of alloying elements. It is thus obvi-
is dictated by the inter-particle spacing, and is simi- ous that both the microhardness under the condition of peak
lar to those (calculated) in some of the liquid phase aging and time required for reaching that depend on the time
sintered Al alloy–SiC composites [14], with coarse and temperature of solution treatment.
particle sizes and low volume fraction, which have Although the rate of increase in hardness with aging time
also shown retarded or unaltered aging kinetics, with varies, yet the time required for peak aging of the composites
respect to that of the Al alloy. The low dislocation appears to be independent of size or volume fraction of SiC
density generated by the smaller thermal expansion particles, which is in agreement with the observations of Pan
mismatch between Al and Al2 O3 has been cited as and Cheng [16], and Suresh et al. [6]. Suresh et al. have shown
one of the reasons for the little effect of reinforcement that once a threshold density of dislocations is reached, further
on aging kinetics of the composite [32]. increase in the SiC volume fraction has little effect on the aging
(b) Static recovery through the annihilation of dislocations kinetics.
of opposite sign is likely to set in and reduce the dislo-
cation density inside the composites at the temperature 4. Conclusions
of aging, which is about half the absolute melting point
of pure Al. Unlike precipitation, annihilation of dislo- The artificial aging behaviour of the powder metallurgy
cations of opposite sign in a dense network reduces the processed Al–Cu–Mg alloy and the matrix in its composites rein-
strain energy of the system, and does not face an acti- forced with 5, 15, or 25 vol.% SiC particles has been studied after
vation barrier. There must be a competition between incomplete or complete solutionizing treatments at 495 ◦ C for
the recovery process and precipitation, with the faster 0.5 h, or 504 ◦ C for 4 h, respectively. The following conclusions
process affecting the hardness more. may be drawn from the present study:
Precipitates of CuAl2 , CuMgAl2 and those enriched in Fe
The present study shows that aging kinetics of the composite have been observed in the alloy matrix as well as at SiC–matrix
with respect to that of the unreinforced alloy may not be always interfaces in the composites in as-extruded and aged conditions.
faster. At the same time, average microhardness of the matrix The CuMgAl2 precipitates appear rod shaped, and aligned along
in the composite with 5 vol.% SiC is the highest for both the the orthogonal 1 0 0 directions. The composite samples solu-
solution treatments, which can be best explained on the basis tion treated at 495 ◦ C for 0.5 h have a larger fraction of the
of the much smaller inter-particle spacing, finer particle size, Al–SiC interfaces with undissolved precipitates and enrichment
and greater dislocation density (Table 1), compared to others. of alloying elements, than those solutionized at 504 ◦ C for 4 h.
Smaller distance between the SiC particles and higher density The time for peak aging of the Al alloy and its composites
of dislocations is expected to reduce spacing between the precip- with 5, 15 or 25 vol.% SiC particles, solution treated at 495 ◦ C
itates, and increase the microhardness of the matrix. In addition, for 0.5 h and aged at 191 ◦ C has been found to be 3 and 5 h,
its lower total area of the particle–matrix interfaces per unit vol- respectively. On the other hand, hardness of the alloy and com-
ume probably reduces the effect of interfacial segregation of the posites solution treated at 504 ◦ C for 4 h has peaked on artificial
alloying elements, increasing the volume fraction of precipitates. aging for 4 and 6 h, respectively. Again higher peak hardness has
Among the samples solution treated at 495 ◦ C for 0.5 h, the been observed on aging of samples after solution treatment at
alloy has shown higher average and peak hardness compared to 504 ◦ C for 4 h, than those aged after solutionizing at 495 ◦ C for
the composites with 15 or 25 vol.% SiC, which is contrary to 0.5 h. The difference in age-hardening behaviour based on the
the expectation. In contrast, on aging after solution treatment at time and temperature of solution treatment has been attributed to
504 ◦ C for 4 h, the alloy has shown lower hardness compared the larger concentration of Cu atoms in solid solution available
to the composites, which is usually expected. The much higher after solution treatment at 504 ◦ C for 4 h, giving rise to a higher
hardness of the alloy, besides faster kinetics of aging in compari- volume fraction of precipitates.
son to composites with 15 or 25 vol.% SiC for solution treatment The composite with 5 vol.% SiC particles has shown the high-
at 495 ◦ C, implies that volume fraction of precipitates is less est hardness, irrespective of the solution treatment condition,
in the latter. Solutionizing at 495 ◦ C was insufficient for com- probably due to the lowest inter-particle spacing, and a higher
plete dissolution of alloying elements segregated at the Al–SiC volume fraction of precipitates among the composites studied.
interfaces. For solution treatment at 495 ◦ C for 0.5 h, the peak-age hardness
Again the less time required for peak aging in case of the of the Al alloy is greater than that of the composites with 15 or
alloys and composites solution treated at 495 ◦ C for 0.5 h com- 25 vol.% SiC particles, due to the lower volume fraction of pre-
S. Pal et al. / Materials Science and Engineering A 480 (2008) 496–505 505

cipitates in case of the latter. On the other hand, the peak-age [13] P. Appendino, C. Badini, F. Marino, A. Tomasi, Mater. Sci. Eng. A 135
hardness of the alloy is less than that of the composites on aging (1991) 275–279.
after solutionizing at 504 ◦ C for 4 h. [14] G.M. Janowski, B.J. Pletka, Metall. Mater. Trans. 26A (1995) 3027–
3035.
The slower kinetics of aging in the matrix of composites is [15] M.K. Jain, V.V. Bhanuprasad, S.V. Kamat, A.B. Pandey, V.K. Varma,
attributed to the lower concentration of vacancies, inadequate B.V.R. Bhat, Y.R. Mahajan, Inter. J. Powder Metall. 29 (3) (1993)
dislocation density, as well as extensive interfacial segregation 267–275.
of alloying elements, such as Cu, Al and Fe, and undissolved [16] Y.M. Pan, H.S. Cheng, in: P.K. Liaw, J.R. Weertman, H.L. Marcus, J.S.
precipitates at the particle–matrix interfaces causing the deple- Santner (Eds.), Morris E. Fine Symposium, TMS, Warrendale, PA, USA,
1991, pp. 143–151.
tion of Cu or Mg in the matrix, and reduction in degree of solute [17] ASTM E112-96, Standard Test Methods for Determining Average Grain
supersaturation inside the matrix. Size, ASTM International, West Conshohocken, PA, USA, 2003.
[18] E.E. Underwood, in: R.T. DeHoff, F.H. Rhines (Eds.), Quantitative
Acknowledgement Microscopy, McGraw-Hill Book Company, New York, 1968, pp. 78–124,
Chapter 4.
[19] L.C. Lovell, F.L. Vogel, J.H. Wernick, Metal Prog. 75 (1959) 96–96D.
The technical assistance rendered by Mr. B.P. Mirdya and [20] W.H. Cubberly, ASM Metals Handbook, 9th ed., vol. 2, Properties and
Mr. Bijon Das with XRD, Mr. D.N. Mahato and Mr. A.K. Jha Selection: Non-Ferrous Alloys and Pure Metals, American Society of Met-
with heat treatment, Dr. R. Maiti and Mr. Sana of SEM, Mr. als, Materials Park, OH, 1979, p. 45.
Ranadhir Bose with TEM, and Mr. B.C. Jena and Mr. B. Deb in [21] R.J. Borg, G.J. Dienes, An Introduction to Solid State Metals, Academic
Metallography Laboratory is gratefully acknowledged. Press, 1988, pp. 1–23, Chapter 1.
[22] S.R. Nutt, R.W. Carpenter, Mater. Sci. Eng. A 75 (1985) 169–177.
[23] G. Liu, Z. Zhang, J.K. Shang, Acta Metall. Mater. 42 (1994) 271–282.
References [24] M.P. Thomas, J.E. King, Scripta Metall. Mater. 31 (2) (1994) 209–214.
[25] G.B. Brook, Precipitation in Metals, Special Report No. 3, Fulmer Research
[1] K.K. Chawla, M. Metzer, J. Mater. Sci. 7 (1972) 34–39. Institute, Stoke Poges, UK, 1963.
[2] R.J. Arsenault, N. Shi, Mater. Sci. Eng. A 81 (1986) 175–187. [26] A.K. Gupta, P. Gaunt, M.C. Chaturvedi, Philos. Mag. 55A (1987) 375–387.
[3] T.G. Nieh, R.F. Karlak, Scripta Metall. 18 (1984) 25–28. [27] C.R. Hutchinson, S.P. Ringer, Metall. Mater. Trans. 31A (2000) 2721–2733.
[4] J.M. Papazian, Metall. Trans. A 19A (1988) 2945–2953. [28] S. Abis, M. Massazza, P. Mengucci, G. Riontino, Scripta Mater. 45 (2001)
[5] T. Christman, S. Suresh, Acta Metall. 36 (7) (1988) 1691–1704. 685–691.
[6] S. Suresh, T. Christman, Y. Sugimura, Scripta Metall. 23 (1989) 1599–1602. [29] N. Gao, L. Davin, S. Wang, A. Cerezo, M.J. Starink, Mater. Sci. Forum
[7] K.K. Chawla, A.H. Esmaeli, A.K. Datye, A.K. Vasudevan, Scripta Metall. 396–402 (2002) 923–928.
Mater. 25 (6) (1991) 1315–1319. [30] K.K. Chawla, Ceramic Matrix Composites, Chapman and Hall, London,
[8] I. Dutta, S.M. Allen, J.L. Hafley, Metall. Trans. A 22A (1992) 2553–2563. 1993, p. 164.
[9] I. Dutta, D.L. Bourell, Mater. Sci. Eng. A 112 (1989) 67–77. [31] G. LeRoy, J.D. Embury, G. Edwards, M.F. Ashby, Acta Metall. 29 (1981)
[10] Jun-shan Lin, Peng-Xing Li, Ren-Jie Wu, Scripta Metall. Mater. 28 (1993) 1509–1522.
281–286. [32] K.-C. Cen, C.-G. Chao, Metall. Mater. Trans. 26A (1995) 1035–1043.
[11] M.P. Thomas, J.E. King, J. Mater. Sci. 29 (1994) 5272–5278. [33] E. Hunt, P.D. Pitcher, P.J. Gregson, in: T. Khan, G. Effenberg (Eds.),
[12] V.K. Varma, Y.R. Mahajan, V.V. Kutumba Rao, Scripta Metall. 37 (4) Advanced Aluminium and Magnesium Alloys, ASM International, PA,
(1997) 485–489. 1991, p. 687.

Vous aimerez peut-être aussi