Sie sind auf Seite 1von 9

Composites Part B 175 (2019) 107140

Contents lists available at ScienceDirect

Composites Part B
journal homepage: www.elsevier.com/locate/compositesb

Effect of Al2O3 nanoparticles content and compaction temperature on


properties of Al–Al2O3 coated Cu nanocomposites
W.S. Barakat a, A. Wagih b, Omayma A. Elkady c, A. Abu-Oqail d, A. Fathy b, *, A. EL-Nikhaily a
a
Mechanical Department., Faculty of Industrial Education, Suez Uni., Egypt
b
Department of Mechanical Design and Production Engineering, Faculty of Engineering, Zagazig University, P.O. Box 44519, Egypt
c
Powder Technology Division, Manufacturing Technology Department, Central Metallurgical R & D Institute, P.O. 87, Helwan, 11421, Cairo, Egypt
d
Mechanical Department, Faculty of Industrial Education, Beni-Suef University, Egypt

A R T I C L E I N F O A B S T R A C T

Keywords: This paper presents experimental investigation on the effect of Al2O3 content and compaction temperature on
Al-Al2O3 coated nanocomposites mechanical and wear properties of Al–Al2O3 coated Cu nanocomposites. With the aim of improving wettability
Microhardness and dispersion of high weight fraction of Al2O3 nanoparticles in Al matrix, Al2O3 nanoparticles were electroless
Wear properties
coated by Cu particles. High-energy ball milling followed by hot compaction was applied to manufacture
Electroless deposition
Al–Al2O3 coated Cu nanocomposites. The results showed that increasing Al2O3 nanoparticles content up to 10 wt
% improved hardness and wear properties of samples compacted at 400 ᵒC, while by increasing Al2O3 content to
15 wt%, the properties were negatively affected due to the agglomeration of Al2O3 nanoparticles and the high
reduction of relative density. However, well dispersion of high content of Al2O3, 15 wt%, was achieved in
samples compacted at 700 MPa which improved hardness by 105% compared to pure Al. The hot compaction
temperature highly influenced structural, mechanical and wear properties of compacted samples. Increasing
compaction temperature, reduced crystallite size (49 nm) and increased the relative density of Al–Al2O3 coated
Cu nanocomposites which improved mechanical and wear properties of the produced samples. Applying elec­
troless deposition of Cu over Al2O3 nanoparticles followed by high-energy ball milling and hot compaction at
700 MPa is an efficient procedure for production of Al–Al2O3 coated Cu nanocomposite with relatively high
Al2O3 content.

1. Introduction for production is the inclusion of Al2O3 particle in liquid or semi-liquid


Al and casting [6–9], at which Al2O3 are entrapped between a high
Processing of metal matrix composites (MMCs) usually requires viscous aluminum metal by the application of centrifugal force. Despite,
mixing of soft metal, high ductility, with hard ceramic material, high its low cost, difficulties in achieving homogenous distribution of Al2O3
strength, to produce a new material with high strength and accepted in Al matrix have been reported especially when Al2O3 particle size
ductility. Recently, these MMCs receive not only researchers’ attention tends to nanosize [10,11]. High-energy ball milling technique has been
but also designers’ attention due to their applicability in many advanced reported to achieve the well dispersion of micro-size Al2O3 particles in
applications. For structural applications, aluminum (Al)-based metal Al matrix [12,13]. However, reducing the particle size to nano-size, the
matrix composites have been widely used in aerospace and automotive dispersion of reinforcement particles became not an easy task using
industry due to their high stiffness, modulus, specific strength and wear conventional mixing due to their large surface-to-volume ratio and lack
resistance [1–4]. of wettability of fine particles in molten Al [14]. To overcome this
Alumina (Al2O3) is one of the widely used particles to reinforce Al problem, several wettability improvement methods have been presented
matrix due to its availability and excellent material properties such as such as mixing ceramic powders with Mg and Ti powder [15] and
high mechanical strength and good chemical stability at elevated tem­ electroless deposition of Ag [16], Ni [17] and Co [18] on the ceramic
peratures [5]. Different methods have been presented in the literature nanoparticles. Using these methods, improved strength and wear rates
for production of A-Al2O3 composites. A simple and economical method of the produced nanocomposites were achieved for relatively ceramic

* Corresponding author.
E-mail address: afmeselhy@zu.edu.eg (A. Fathy).

https://doi.org/10.1016/j.compositesb.2019.107140
Received 4 April 2019; Received in revised form 20 June 2019; Accepted 5 July 2019
Available online 9 July 2019
1359-8368/© 2019 Elsevier Ltd. All rights reserved.
W.S. Barakat et al. Composites Part B 175 (2019) 107140

small volume fraction [16,17]. gas for 2 h to avoid formation of undesired phases.
Increasing reinforcement volume fraction improves mechanical and Mixtures of Al with 5, 10, 15 wt% Al2O3 coated were prepared using
wear properties of the produced nanocomposites [19–21]. However, this a ball mill machine for 6 h. The ball to powder weight ratio was kept at
rule of thumb is achievable up to a threshold at which agglomeration 1:4. The mixing process was carried out at room temperature and the
and poor dispersion of ceramic nanoparticles occur [22]. Thus, most of rotation speed of the ball mill was kept at 300 rpm. To avoid oxidation of
the available research on metal matrix nanocomposite focusing on small Al, the mixing vial was filled with argon gas before mixing. After that the
volume fraction reinforcement. For example, Kang et al. [22] studied the mixed powders were consolidated by hot pressing at two different
influence of micro and nano Al2O3 particles addition of the properties of temperatures 400 � C and 600 � C with heating rate 10 � C/min under
Al–Al2O3 nanocomposites. Their results show a noticeable improvement 700 MPa pressure for 45 min dwell time under an argon atmosphere.
of strength with the addition of Al2O3 nanoparticle with low volume The die used for consolidation was stainless steel cylinder with 12 mm
fraction up to 4%. However, increasing the volume fraction more than internal diameter.
this value, strength reduction occurs due to Al2O3 nanoparticles clus­
tering. This threshold volume fraction can be increased by electroless 2.2. Composite characterization
deposition of different metals that have better thermal stability and good
wettability with Al such as Cu. Densification parameters of sintered samples in terms of apparent
The main objective of this work is to present a novel method for porosity and bulk density were measured in an aqueous media by
production of Al–Al2O3 nanocomposite with relatively high-volume Archimedes principle, using water as a floating liquid according to JIS
fraction using electroless deposition of Cu nanoparticles over Al2O3, R2205-1992.
high energy ball milling and hot compaction. Moreover, the effect of The sintered samples were grinded using grit papers with grades 500
Al2O3 coated nanoparticles content on the mechanical and wear prop­ up to 2500 and then polishing by 6 μm and 1 μm alumina paste and
erties of the produced nanocomposite is investigated. Additionally, the etched with Keller’s reagent for microstructural analysis. The micro­
influence of hot compaction temperature on the microstructure, me­ structure of the polished samples was investigated by field emission
chanical and wear properties of the produced nanocomposite and its scanning electron microscope (FE-SEM).
effect on the threshold content of Al2O3 coated are investigated. The phase formation and composition changes during the sintered
process were identified by X-ray diffraction analysis using X-ray
2. Material and experiments diffractometer with an automated Siemens Model D-5000 diffractom­
eter using Cu K-alpha radiation and operated at 40 KV and 30 mA in the
2.1. Materials preparation 2θ range of 20� –100� by the step of 0.05� .
The hardness of the consolidated samples was measured by Vickers
Samples used in this experiment were made of commercial Al pow­ microhardness tester at room temperature, on the polished surface. Each
der of 99.99% purity with particle size less than 10 μm (Dop organic value was an average of at least six randomly hardness indentations by
kimya, Ankara, Turkey). Al2O3 powder (M K Impex CORP. Canada.) applying 300 g load with loading time of 15 s.
with 99.99% purity and 40 nm particle size was used as a reinforcement. Dry sliding wear rate of all samples was carried out using a pin-on-
The electroless deposition process is analogous to electroplating with a disc at room temperature without lubrication. The pins of 64 mm2 in
slight difference that electroless deposition does not need electrical contact area for all samples were tested by sliding them on a hardened
currents. In electroless deposition, the electron transferring takes place steel disc with hardness of 64 HRC and surface roughness of 60 μm under
through chemical reduction process leading to deposition of metal on different loads. The experiments were carried out under four different
the surface of ceramic by reducing agent. Before the deposition of Cu on normal loads, 5, 10, 15 and 20 N for 1500 rotations, corresponding to a
Al2O3, pretreatment of Al2O3 for favorable condition for plating was fixed sliding distance of 471 m, at constant sliding speeds of 1.5 m/s.
applied. The pretreatment of Al2O3 was performed in two steps. The first During sliding, the wear rate was defined as the weight loss suffered per
step was sensitization of Al2O3 nanoparticles by immersing on 10 wt% unit sliding distance. To compute the wear rate, the wear pin was
sodium hydroxide solution with magnetic stirring for 1 h, then in cleaned by in acetone and weighed on a microbalance with �0.1 mg
acetone with stirring for 1 h with the aim of cleaning their surface to sensitiveness before and after the wear tests. The coefficient of friction
eliminate any impurities of ceramic crystals. After that the powders were was continuously recorded during wear test by measured the frictional
filtered and washed by distilled water then dried in electrical furnace at force using a load cell then divided by the normal load on the specimen.
110 � C for 1 h.
The second step was activation of Al2O3 nanoparticles for surface 3. Results
metallization. This step is mandatory for such non-conducting materials,
Al2O3 particles, in which the direct deposition of cu is impossible. 3.1. Microstructure of powder
Therefore, activation for Al2O3 nanoparticles surfaces was performed by
deposition of silver on their surfaces to facilitate attraction of copper Fig. 1 shows microstructure of Al2O3 nanoparticles before and after
particles during electroless deposition process. Chemical bath of 3 g/l electroless coating by Cu. As shown in the figure, the microstructure of
silver nitrate and 300 ml formaldehyde was prepared with PH within the as received powder is almost the same as the powder after coating.
range 11–13 was used for deposition of 10 wt% silver on Al2O3 nano­ The powders appear after coating with larger dispersion and less
particles surface. After silver chemical depositions, the solution was agglomeration which facilitate the dispersion of Al2O3 nanoparticle in Al
filtrated, washed using distilled water and dried in electrical furnace at matrix during ball milling. Fig. 1(c) shows the EDX analysis of the coated
110 � C for 1 h. The electroless deposition of 30 wt% Cu on Al2O3 powder which reflect the presence of Cu, Ag and Al2O3 only without the
nanoparticles was performed using chemical path of 35 g/l copper sul­ accumulation of any other materials during coating process. Fig. 2 shows
fate pentahydrate (CuSO4.5H2O), 200 ml/l formaldehyde reducing XRD analysis of Al2O3 before and after coating with Ag and Cu. The
agent and 170 g/l Rochelle salt (C4H4O6KNa.4H2O). Rochelle salt was peaks are shown for only the three phases which also confirm the val­
added to the solution because copper sulfate was not soluble in PH above idity of the electroless coating process for coating Al2O3 nanoparticles
4 which increase the complexity of the deposition process. The pH de­ with Cu particles without contamination of other undesired chemicals.
gree of the solution was adjusted greater than 11 using sodium hy­
droxide (NaOH). Then, the solution was filtrated, washed by distilled 3.2. Microstructure of consolidated samples
water and dried in electrical furnace at 110 � C for 1 h. The coated
nanoparticles were heated up to 500 � C on tube furnace under hydrogen Figs. 3 and 4 show microstructures of the consolidated samples at

2
W.S. Barakat et al. Composites Part B 175 (2019) 107140

Fig. 1. Microstructure of Al2O3 nanoparticles (a) before coating, (b) after coating and (c) EDX of powder after coating.

Fig. 3. SEM micrographs of hot compacted samples at 400 � C: (a) Al, (b) Al–5%
Fig. 2. XRD of Al2O3 nanoparticles after Ag and Cu coating. Al2O3, (b) Al–10%Al2O3 and (b) Al–15%Al2O3.

two different hot compaction temperatures, 400 ᵒC and 600 ᵒC, respec­ the sample and facilitate the tendency to form macro-sized voids as
tively. In Fig. 3, micro voids are observed in all the consolidated samples shown in Fig. 3(d) for sample contain 15 wt% Al2O3. This due to the
even in the one that is free of reinforcement (see Fig. 3(a)). Reinforcing higher compressive strength of Al2O3 coated nanoparticles compared to
metallic material with ceramics always leads to formation of pores due Al which makes the required pressure for consolidating Al at 400 ᵒC is
to the large difference between surface energy of ceramics and metals quite low for consolidating Al2O3. In Fig. 4, samples consolidated at 600
[13,23–25]. Also, the arbitrary shape of ceramic phase influences the ᵒ
C with different Al2O3 coated content, 0, 5, 10 and 15 wt% are shown. It
amount of voids in the composite during compaction [26]. Increasing is observed that the micro-voids are less presented in these samples than
coated Al2O3 nanoparticles content increases the micro void density in that compacted at 400 ᵒC. The distribution of Al2O3 coated nanoparticles

3
W.S. Barakat et al. Composites Part B 175 (2019) 107140

is homogeneous in all the samples. To ensure the homogeneity of the


consolidated micro-structure, mapping analysis for the sample con­
taining the highest Al2O3 content is shown in Fig. 5. The figure shows the
homogeneous distribution of the constituents without any observation
of agglomeration.
Fig. 6 shows XRD analysis of Al–Al2O3 nanocomposite with different
content of Al2O3 coated Cu nanoparticles compacted at 400 and 600 ᵒC.
The figure confirms the composition of the prepared nanocomposite
without any evidence of precipitation of other undesired chemical
during deposition of Ag and Cu particles. Peak broadening is observed
with increasing Al2O3 content for samples compacted at 400 and 600 ᵒC.
For samples compacted at 400 ᵒC, there is no evidence on reaction be­
tween Cu and Al forming intermetallic phases, while in sample contain
15 wt% Al2O3 and compacted at 600 ᵒC, peaks for intermetallic phase
appear which indicate that reaction occur between Al and Cu. This
intermetallic phase negatively influences the electrical and thermal
properties of the produced nanocomposite [26]. The crystallite size of
the prepared samples with different coated Al2O3 and consolidated at
Fig. 4. SEM micrographs of hot compacted samples at 600 � C: (a) Al, (b) Al–5% different temperature is shown in Table 1. It is observed that the crys­
Al2O3, (b) Al–10%Al2O3 and (b) Al–15%Al2O3.
tallite size is reducing with increasing coated Al2O3 content. Addition­
ally, it decreases with increasing the consolidating temperature.

Fig. 5. Mapping of Al–Al2O3 nanocomposite compacted at 600 � C.

4
W.S. Barakat et al. Composites Part B 175 (2019) 107140

Fig. 7. Microhardness of Al–Al2O3 coated Cu nanocomposites consolidated at


two different temperatures: (a) 400 � C and (b) 600 � C.

Fig. 6. XRD analysis of Al–Al2O3 coated Cu nanocomposite consolidated at two


different temperatures: (a) 400 � C and (b) 600 � C.

3.3. Mechanical and wear properties

The microhardness of the Al–Al2O3 coated Cu nanocomposites


consolidated at 400 and 600 ᵒC is shown in Fig. 7. It is observed that
increasing coated Al2O3 nanoparticles content up to 10 wt% improves
microhardness of the consolidated samples at 400 ᵒC. However, the
microhardness is reduced by increasing coated Al2O3 content to 15 wt%.
Unlike, in the samples consolidated at 600 ᵒC, the hardness almost lin­
early increases with increasing coated Al2O3 content.
Fig. 8 shows the wear rate of Al–Al2O3 coated Cu nanocomposites at
different loads, 5, 10, 15 and 20 N and consolidated at different tem­
peratures, 400 and 600 ᵒC. Generally, increasing the applied load in­
creases the wear rate due to the higher penetration of indenter in the
sample which enable higher metal removal rate [27,28]. For samples
consolidated at 400 ᵒC, at the same applied load, the wear rate is reduced
with increasing coated Al2O3 nanoparticles content up to 10 wt%. Fig. 8. Wear rate of Al–Al2O3 coated Cu nanocomposites consolidated at two
However, for larger coated Al2O3 content, the wear rate is increased different temperatures: (a) 400 � C and (b) 600 � C.
again. This is not the case for the samples consolidated at 600 ᵒC at
which the wear rate is reduced with increasing coated Al2O3 content The average coefficient of friction of Al–Al2O3 coated Cu nano­
even for sample contain 15 wt% Al2O3. This observation highlights large composites at different loads, 5, 10, 15 and 20 N and consolidated at two
dependence of wear properties of Al–Al2O3 coated Cu nanocomposites different temperatures, 400 and 600 ᵒC is shown in Fig. 9. The coefficient
on the compaction temperature. of friction is reduced with increasing load due to the higher indentation
depth of the pin in the sample. It is observed that increasing coated
Al2O3 content up to 10 wt% reduces the coefficient of friction of samples
Table 1
compacted at 400 ᵒC. However, increasing coated Al2O3 content to 15 wt
Crystallite size of Al–Al2O3 coated Cu nanocomposites.
% negatively affect the coefficient of friction of the produced composite.
Al2O3% Crystallite size (nm) For samples compacted at 600 ᵒC, the coefficient of friction reduces with
400 � C 600 � C increasing coated Al2O3 due to the presence of Al2O3 hard particles in
0 157.07 290.34
the structure. Again, the compaction temperature influences the prop­
5 143.52 55.90 erties of the manufactured nanocomposites especially at high coated
10 138.63 53.73 Al2O3 content.
15 116.46 49.02 Based on the obtained results and observations, it is believed that

5
W.S. Barakat et al. Composites Part B 175 (2019) 107140

Al2O3 and compaction temperature influence the mechanical and properties of Al2O3 nanoparticles compared to Al which make the
tribological properties of Al–Al2O3 coated Cu nanocomposites. In the consolidation of these composites harder than the pure metal. Addi­
following section, the obtained results are discussed to highlight the tionally, the irregular shape of Al2O3 nanoparticles with sharp edges
influence of coated Al2O3 content and compaction temperature on results entrapping air between particles acting as pores. The relative
properties of the produced nanocomposites. Also, the interplay between density decreasing rate is almost constant with increasing Al2O3 content
microstructural, mechanical and tribological properties is discussed. reaching 96.13% for sample contain 15 wt% Al2O3 nanoparticles which
indicates that the consolidation process is almost the same for all the
4. Discussion prepared composites (see Fig. 11).
The crystallite size is reducing with increasing Al2O3 content as
4.1. Effect of Al2O3 coated Cu content shown in Table 1. During high energy ball milling, two main mecha­
nisms occurs: initially, welding of Al particles entrapping between them
Generally, increasing reinforcement volume content improves Al2O3 nanoparticles and fragmentation of the welded particles at longer
microstructural, mechanical and tribological properties if a homogenous milling times [12,29]. So, during such experiment, two main reasons for
distribution of the reinforcement is achieved [3,13,16,29]. So, homog­ crystallite size reduction: the fragmentation of Al particles and the
enous distribution of reinforcement phase in the matrix is a key entrapping of Al2O3 nanoparticles between Al particles which act as
parameter for improving properties of the obtained nanocomposites. impedance for dislocation movement. The presence of coated Al2O3
Another important parameter that influence properties of nano­ nanoparticles during high energy ball milling of Al particles results in
composites is the voids in the sample which give indication on the generation of dislocations and thus, the crystallite size is reduced [29,
density of the produced samples. For samples compacted at 400 ᵒC, 30]. Additionally, during compaction at high temperature, the crystal­
increasing coated Al2O3 content, increases the micro-voids in the sample lite size can be reduced as well due to the recrystallization of grains at
(see Fig. 3) until formation of macro-voids in sample contain 15 wt% high temperature [31].
coated Al2O3. In order to quantify the influence of coated Al2O3 addition Microhardness of Al–Al2O3 coated Cu nanocomposites is observed to
on the pores formation, the relative density versus Al2O3 content is be improved with increasing Al2O3 content up to 10 wt% for samples
shown in Fig. 10. As shown in the figure, the relative density is decreased compacted at 400 ᵒC despite the reduction in the relative density. It
to 94.1% for sample contain 10 wt% Al2O3 and consolidated at 400 ᵒC increases from 47.5 HV for Al to 74.2 HV for Al-10 wt%Al2O3 coated Cu
while it reduces to 89.4% for sample contain 15 wt% Al2O3 which reflect nanocomposite achieving 56.2% increase. In these composites, two pa­
larger decreasing rate of relative density for this sample. This is due to rameters affect the microhardness in an opposite manner. The reduction
the agglomeration of coated Al2O3 nanoparticles during the consolida­ of relative density negatively affects the general mechanical properties.
tion process (see Fig. 3(d)). The agglomeration of these particles in an However, the grain refinement and dislocation generation strengthen
irregular shape results in more voids between Al and Al2O3 particles the composite. In these composites, the positive effect of grain refine­
[29]. For samples consolidated at 600 ᵒC, the relative density of sample ment and dislocation generation is able to vanish the negative effect of
free of coated Al2O3 nanoparticles is almost 100% which mean that the relative density reduction. Therefore, this improvement in the micro­
prepared sample is fully dense. Addition of 5 wt% coated Al2O3 de­ hardness is attributed to the grain refinement/crystallite size reduction
creases the relative density to 98.84% due to the higher compressive (see Table 1). Additionally, the higher microhardness of Al2O3 nano­
particles and their coating layer (Cu nanoparticles) contribute this
improvement. Increasing Al2O3 content to 15 wt% results in a slight
reduction of the microhardness of the prepared composite due to the
agglomeration of Al2O3 nanoparticles (see Fig. 3(d)) and the larger
reduction of the relative density (see Fig. 10). So, in this sample, two
parameters negatively affect the mechanical properties against slight
grain refinement and crystallite size reduction (see Table .1). For the
samples compacted at 600 ᵒC, the microhardness is increased with Al2O3
content from 58.5 for Al to 120.4 for sample contain 15 wt% achieving
105% improvement. This large increase due to the higher grain refine­
ment and crystallite size reduction (see Table 1) besides the higher
hardness of Al2O3 and Cu particles as previously stated.

Fig. 9. Coefficient of friction of Al–Al2O3 coated Cu nanocomposites consoli­ Fig. 10. Effect of Al2O3 content on relative density of Al–Al2O3 coated Cu
dated at two different temperatures: (a) 400 � C and (b) 600 � C. nanocomposites.

6
W.S. Barakat et al. Composites Part B 175 (2019) 107140

facilitate the material removal process [40]. The coefficient of friction of


the prepared samples during abrasive wear test follow almost the same
trend observed for wear rates (see Fig. 9).

4.2. Effect of compaction temperature

Hot compaction process is presented in this work to consolidate the


milled powder instead of the conventional consolidation process, cold
compaction followed by sintering. Hot compaction process has advan­
tages over the conventional consolidation procedure such as shorter
processing time and less possibility for void formation. The compaction
temperature seems to highly influence physical, mechanical and wear
properties of Al–Al2O3 coated Cu nanocomposites. Increasing hot
compaction temperature provides more dense samples with less amount
Fig. 11. Comparison between hardness improvement in the current samples of pores. It is observed in Fig. 10 that for the samples with the same
consolidated at 600 � C and available results in the literature.
Al2O3 content, samples compacted at 600 ᵒC show larger relative density
than those compacted at 400 ᵒC. For example, the relative density of Al-
In order to highlight the effect of Cu coating on the mechanical 10 wt%Al2O3 coated Cu nanocomposite compacted at 600 ᵒC is 3.1%
properties, the microhardness of the samples consolidated at 600 � C was larger than that compacted at 400 ᵒC. At higher compaction tempera­
compared with the properties of the same composite available in the ture, near Al melting temperature, plastic deformation of grains occurs
literature [9,13,22,32–34] without coating Al2O3. The figure represents due to thermal stresses and consequently, the material flows easily to fill
the improvement factor of microhardness by normalizing the micro­ the pores. Fig. 12 shows SEM micrograph of interface between four
hardness of the produced nanocomposites with the microhardness of the grains and three grains in Al-10 wt%Al2O3 coated Cu nanocomposite
raw Al. It is observed that the microhardness of Al–Al2O3 nano­ compacted at 400 and 600 ᵒC, respectively. As shown in the figure,
composites improves with increasing Al2O3 content up to a certain value micro-pores are observed at the interface due to the low flowability of
for all the compared samples. For Kang et al. [22] and Ezatpour et al. the metal during hot compaction at 400 ᵒC. However, there is no evi­
[33] results, the maximum improvement is achieved at 5 wt% Al2O3, dence on formation of micropores between grains in the sample com­
while further increasing of the Al2O3 content results in microhardness pacted at 600 ᵒC. Additionally, Al2O3 particles are found in small
reduction. For the other materials shown in the figure (Wagih and Fathy agglomeration in sample compacted at 400 ᵒC (see Fig. 12(a)).
[9], Wagih [13], Mahbob et al. [32], Sadeghi et al. [34]), increasing The crystallite size of the hot compacted samples is influenced by the
Al2O3 content more than 5%, the microhardness improvement rate is compaction temperature as shown in Table 1. For pure Al, samples
reduced. Quantitively, the prepared composites in this study with 10 compacted at 600 ᵒC show larger crystallite size than those compacted at
and 15 wt% Al2O3 coted Cu show the maximum microhardness among 400 ᵒC due to the recrystallization and grain growth occur during hot
the available results in the literature. Interestingly, the microhardness compaction. Compaction of Al at 600 ᵒC, near the melting point, allows
improvement rate in the current study is constant up to 15 wt% Al2O3 the grain recrystallization and reduces the lattice strain due to grain
which indicate that the microhardness can be improved by increasing relaxation [31]. For samples contain Al2O3 nanoparticles, opposite trend
reinforcement weight fraction. This large and continuous improvement is observed, samples compacted at 600 ᵒC shows smaller crystallite size,
in the current study is attributed to the coating of Al2O3 with Cu due to the presence of Al2O3 on the grain boundaries which creates
nanoparticles which improve the surface properties, wettability and crystal defects such as dislocation buildup and prevents dislocation
dispersion of Al2O3 nanoparticles in Al matrix (see Figs. 1, 4 and 5). movement and grain growth [13,41,42].
The wear rate and coefficient of friction are highly influenced by the Microhardness of pure Al and Al–Al2O3 coated Cu nanocomposites is
addition of coated Al2O3 nanoparticles as shown in Figs. 8 and 9. dependent on the compaction temperature as shown in Fig. 7. For pure
Generally, addition of Al2O3 nanoparticles reduces the wear rate for all Al, microhardness of samples compacted at 600 ᵒC is larger than for
samples which indicates less material removal rate while the SiC parti­ samples compacted at 400 ᵒC, although, the crystallite size is larger. The
cles penetrate the surface of the prepared nanocomposites. The presence mechanical properties of compacted samples depend on the grain size of
of Al2O3 coated Cu nanoparticles increases the impedance of the mate­ the compacted material and the relative density [25,30]. The presence
rial for SiC particles penetration which reduces the material removal of pores inside the material, reduces its resistance for penetration and
rate. Moreover, the improved hardness of Al–Al2O3 nanocomposites hence the hardness is reduced. So, for pure Al samples, the negative
than Al contributes the reduction of wear rate according to Archard law effect of lower relative density on the hardness is dominant and over­
[16,35–37]. For samples consolidated at 400 ᵒC, the major impact of come the positive effect of smaller crystallite size. Unlike, samples
Al2O3 is observed at 5 wt%. A slight impact of Al2O3 content is observed contain coated Al2O3 nanoparticles, the higher compaction temperature
by increasing its content to 10 wt%. However, by increasing Al2O3
content to 15 wt%, the wear rate is increased again due to the agglom­
eration of Al2O3 nanoparticles in this composite (see Fig. 3 (d)). Addi­
tionally, the decrease of hardness of this material (see Fig. 7) contributes
the higher wear rate than sample contain 10 wt% Al2O3. This behavior
disappears in samples compacted at 600 ᵒC, at which, the wear rate is
reduced with increasing Al2O3 content. This result correlates well with
the hardness measurement (see Fig. 7) in which, the hardness is pro­
portional to Al2O3 content. The applied load during wear test influences
the wear rate as well. Increasing the applied load results in increased
wear rate because the penetration of SiC particle in the sample can be
simulated as indentation problem [3,29,38,39], at which increasing the
applied load results in larger indentation depth [39]. When the inden­
tation depth becomes larger than the critical indentation depth, large Fig. 12. SEM micrograph of Al–10%Al2O3 coated Cu nanocomposite consoli­
debris are removed causing more plastic deformation of subsurface and dated at two (a) 400 � C and (b) 600 � C.

7
W.S. Barakat et al. Composites Part B 175 (2019) 107140

is, the larger microhardness is. For example, samples compacted at 600 fracture process and generate crystal defects. The relative density was

C shows 24.9%, 38.2% and 71.7% larger hardness than samples contain reduced by increasing Al2O3 content due to the formation of micropores
5, 10 and 15 wt% Al2O3 coated Cu nanoparticles compacted at 400 ᵒC, during consolidation process. Homogeneous dispersion of coated Al2O3
respectively. This due to the smaller crystallite size of samples com­ nanoparticles in Al matrix was achieved up to 10 wt% for sample
pacted at 600 ᵒC (see Table 1) and the higher relative density (see consolidated at 400 ᵒC, however increasing Al2O3 content to 15 wt%,
Fig. 10). Interestingly, for the sample with the largest Al2O3 content, large particle agglomeration was occurred. For samples compacted at
15 wt% and compacted at 600 ᵒC, microhardness is still improving 600 ᵒC, good dispersion was achieved for relatively high Al2O3 content.
which is not the case of the sample compacted at 400 ᵒC due to the Microhardness increased with increasing Al2O3 achieving 105%
prohibition of agglomeration (see Fig. 4) and the reduction of voids (see improvement for samples contain 15 wt% Al2O3 and compacted at 600

Fig. 10). The prohibition of agglomeration and the better density of C due to the crystallite size reduction and well dispersion of Al2O3
samples compacted at 600 ᵒC results in better bonding between grains nanoparticles. The improvement in hardness and crystallite size reduc­
which reduces the ability of the subsurface for plastic deformation tion improved the wear properties of Al–Al2O3 coated Cu
during indentation event [43,44]. nanocomposites.
The wear rate of Al–Al2O3 coated Cu nanocomposites compacted at The properties of Al–Al2O3 coated Cu nanocomposites were highly
400 ᵒC and 600 ᵒC and measured at two different test load, 5 and 15 N is influenced by hot compaction temperature. The crystallite size of pure
replotted in Fig. 13 to highlight the effect of compaction temperature on Al was increased by increasing hot compaction temperature due to the
wear rates. Increasing hot compaction temperature, reduce the material recrystallization and relaxation of grain caused by high temperature,
removal rate during wear test due to the better distribution of coated while, it decreased for Al–Al2O3 coated Cu nanocomposites. The
Al2O3 nanoparticles and reduction of voids (see Fig. 12). reduction of crystallite size was because of the presence of Al2O3
Based on the presented results and discussion, the hot compaction nanoparticles on the grain boundaries which prevent the grains relax­
temperature has large influence on the structural, physical, mechanical ation and recrystallization. Increasing compaction temperature to 600 ᵒC
and wear properties of Al–Al2O3 coated Cu nanocomposites. Increasing helped for achieving nanocomposite with higher Al2O3 content reaching
compaction temperature results in better dispersion of Al2O3 nano­ 15 wt% without agglomeration which improves hardness by 105% and
particles in Al matrix which provide better impedance of dislocation reduce the wear rate by 52% compared to pure Al.
movement and hence less crystallite size. Additionally, this high tem­
perature allows higher flowability of the metal during hot compaction References
process which reduces the probability to form voids and entrapping air.
The less void content provides better bonding between grains and [1] Fathy A, Sadoun A, Abdelhameed M. Effect of matrix/reinforcement particle size
ratio (PSR) on the mechanical properties of extruded Al–SiC composites. Int J Adv
therefore less ability for plastic deformation which results in improved Manuf Technol 2014;73:1049–56.
hardness and wear properties. [2] Fathy A, Ibrahim D, Elkady O, Hassan M. Evaluation of mechanical properties of
1050-Al reinforced with SiC particles via accumulative roll bonding process.
J Compos Mater 2019;53(2):209–18.
5. Conclusion [3] Wagih A, Fathy A. Experimental investigation and FE simulation of spherical
indentation on nano-alumina reinforced copper-matrix composite produced by
Electroless deposition of Cu on Al2O3 nanoparticles followed by high- three different techniques. Adv Powder Technol 2017;28:1954–65.
[4] Fathy A, Wagih A, Abu-Oqail A. Effect of ZrO2 content on properties of Cu-ZrO2
energy ball milling and hot compaction was applied to manufacture nanocomposites synthesized by optimized high energy ball milling. Ceram Int
Al–Al2O3 coated Cu nanocomposites. Four different Al2O3 weight frac­ 2019;45(2):2319–29.
tions, 0, 5, 10 and 15 wt%, were considered to study the influence of [5] Chen WH, Lin HT, Nayak PK, Huang JL. Material properties of tungsten
carbide–alumina composites fabricated by spark plasma sintering. Ceram Int 2014;
Al2O3 coated Cu nanoparticles addition on the properties of Al–Al2O3
40(9):15007–12.
nanocomposite. Additionally, hot compaction was applied at two [6] Roshan M, Mousavian TR, Ebrahimkhani H, Mosleh A. Fabrication of Al-based
different temperatures, 400 ᵒC and 600 ᵒC to investigate the compaction composites reinforced with Al2O3–Tib2 ceramic composite particulates using
temperature effect on the properties of the sample. Structural, me­ vortex-casting method. J Min Metall Sect B Metall 2013;49:299–305.
[7] Mousavian RT, Damadi S, Khosroshahi RA, Brabazon D, Mohammadpour M.
chanical and wear properties of Al–Al2O3 coated Cu nanocomposite A comparison study of applying metallic coating on SiC particles for manufacturing
were investigated. of cast aluminum matrix composites. Int J Adv Manuf Technol 2015:1–12.
Increasing Al2O3 coated Cu nanoparticles content in Al matrix [8] Kumar S, Panwar RS, Pandey O. Effect of dual reinforced ceramic particles on high
temperature tribological properties of aluminum composites. Ceram Int 2013;39:
resulted in reduction of crystallite size due to the entrapping of these 6333–42.
particles between Al particles during ball milling which facilitate [9] Wagih A, Fathy A. Experimental investigation and FE simulation of nano-
indentation on Al–Al2O3 nanocomposites. Adv Powder Technol 2016;27(2):
403–10.
[10] Aguilar J, Fehlbier M, Ludwig A, Bührig-Polaczek A, Sahm P. Non-equilibrium
globular microstructure suitable for semisolid casting of light metal alloys by rapid
slug cooling technology (RSCT). Mater Sci Eng A 2004;375:651–5.
[11] Nafisi S, Ghomashchi R. Effect of stirring on solidification pattern and alloy
distribution during semi-solid-metal casting. Mater Sci Eng A 2006;437:388–95.
[12] Fathy A, Wagih A, El-Hamid MA, Hassan A. Effect of mechanical milling on the
morphology and structural evaluation of Al-Al2O3 nanocomposite powders. Int J
Eng Trans A 2013;27:625–32.
[13] Wagih A. Synthesis of nanocrystalline Al2O3 reinforced Al nanocomposites by high
energy mechanical alloying: microstructural evolution and mechanical properties.
Trans Indian Inst Met 2016;69(4):851–7.
[14] Wagih A, Fathy A. Improving compressibility and thermal properties of Al–Al2O3
nanocomposites using Mg particles. J Mater Sci 2018;53(16):11393–402 [].
[15] Fathy A, Wagih A, El-Hamid MA, Hassan AA. The effect of Mg add on morphology
and mechanical properties of Al–xMg/10Al2O3 nanocomposite produced by
mechanical alloying. Adv Powder Technol 2014;25(4):1345–50.
[16] Wagih A, Abu-Oqail A, Fathy A. Effect of GNPs content on thermal and mechanical
properties of a novel hybrid Cu-Al2O3/GNPs coated Ag nanocomposite. Ceram Int
2019;45:1115–24.
[17] Khosroshahi NB, Khosroshahi RA, Mousavian RT, Brabazon D. Effect of electroless
coating parameters and ceramic particle size on fabrication of a uniform Ni–P
Fig. 13. Effect of hot compaction temperature on wear rate of Al–Al2O3 coated coating on SiC particles. Ceram Int 2014;40:12149–59.
Cu nanocomposite. In the legend, 5 N and 15 N refer to the load during
wear test.

8
W.S. Barakat et al. Composites Part B 175 (2019) 107140

[18] Abu-Oqail A, Samir A, Essa ARS, Wagih A, Fathy A. Effect of GNPs coated Ag on [32] Mahboob H, Sajjadi SA, Zebarjad SM. Influence of nanosized Al2O3 weight
microstructure and mechanical properties of Cu-Fe dual-matrix nanocomposite. percentage on microstructure and mechanical properties of Al–matrix
J Alloy Comp 2019;781:64–74. nanocomposite. Powder Metall 2011;54(2):148–52.
[19] Gudlur P, Boczek A, Radovic M, Muliana A. On characterizing the mechanical [33] Ezatpour HR, Torabi-Parizi M, Sajjadi SA. Microstructure and mechanical
properties of aluminum–alumina composites. Mater. Sci. Eng. A 2014;590:352–9. properties of extruded Al/Al2O3 composites fabricated by stir-casting process.
[20] Sohi MH, Hojjatzadeh SMH, Moosavifar SS, Heshmati-Manesh S. Liquid phase Trans Nonferrous Metals Soc China 2013;23(5):1262–8.
surface melting of AA8011 aluminum alloy by addition of Al/Al2O3 nano- [34] Sadeghi B, Rizzo A, Cavaliere P, Ashrafizadeh F, Shamanian M. Wear behavior of
composite powders synthesized by high-energy milling. Appl Surf Sci 2014;313: Al-based nanocomposites reinforced with bimodal micro-and nano-sized Al2O3
76–84. particles produced by spark plasma sintering. Mater Perform Char 2018;7(1):
[21] Hossein-Zadeh M, Mirzaee O, Saidi P. Structural and mechanical characterization 327–50.
of Al-based composite reinforced with heat treated Al2O3 particles. Mater Des [35] Alpas AT, Zhang J. Effect of microstructure (particulate size and volume fraction)
2014;54:245–50. 1980-2015. and counterface material on the sliding wear resistance of particulate-reinforced
[22] Kang YC, Chan SLI. Tensile properties of nanometric Al2O3 particulate-reinforced aluminum matrix composites. Metall Mater Trans A 1994;25:969–83.
aluminum matrix composites. Mater Chem Phys 2004;85(2–3):438–43. [36] Fathy A, Elkady O, Abu-Oqail A. Microstructure, mechanical and wear properties
[23] Wagih A, Fathy A, Kabeel AM. Optimum milling parameters for production of of Cu–ZrO2 nanocomposites. Mater Sci Technol 2017;33:2138–46.
highly uniform metal-matrix nanocomposites with improved mechanical [37] Fathy A, Shehata F, Abdelhameed M, Elmahdy M. Compressive and wear resistance
properties. Adv Powder Technol 2018;29(10):2527–37. of nanometric alumina reinforced copper matrix composites. Mater Des 2012;36:
[24] Abu-Oqail A, Wagih A, Fathy A, Elkady O, Kabeel AM. Effect of high energy ball 100–7.
milling on strengthening of Cu-ZrO2 nanocomposites. Ceram Int 2019;45(5): [38] Fischer-Cripps A. In: Nanoindentation. third ed. New York: Springer; 2011.
5866–75. [39] Wagih A, Maimi P, Blanco N, Trias D. Predictive model for the spherical
[25] Wagih A, Fathy A, Sebaey TA. Experimental investigation on the compressibility of indentation of composite laminates with finite thickness. Compos Struct 2016;153:
Al/Al2O3 nanocomposites. Int J Mater Prod Technol 2016;52:312–32. 468–77.
[26] Tsuzuki A, Sago S, Hirano SI, Naka S. High temperature and pressure preparation [40] Tjong SC, Lau KC. Abrasive behavior of TiB2 particle-reinforced copper matrix
and properties of iron carbides Fe7C3 and Fe3C. J Mater Sci 1984;19(8):2513–8. composites. Mater Sci Eng A 2000;282:183–6.
[27] Fathy A, Abu-Oqail A, Wagih A. Improved mechanical and wear properties of [41] Wagih A. Mechanical properties of Al-Mg/Al2O3 nanocomposite powder produced
hybrid Al-Al2O3/GNPs electro-less coated Ni nanocomposite Ceramics. by mechanical alloying. Adv Powder Technol 2015;26(1):253–8.
International 2018;44:22135–45. [42] Zhang Z, Chen DL. Consideration of Orowan strengthening effect in particulate-
[28] Tjong SC, Lau KC. Tribological behavior of SiC particle-reinforced copper matrix reinforced metal matrix nanocomposites: a model for predicting their yield
composites. Mater Lett 2000;43(5):274–80. strength. Scripta Mater 2006;54(7):1321.
[29] Suryanarayana C, Al-Aqeeli N. Mechanically alloyed nanocomposites. Prog Mater [43] Wagih A, Fathy A, Ibrahim D, Elkady O, Hassan M. Experimental investigation on
Sci 2013;58(4):383–502. strengthening mechanisms in Al-SiC nanocomposites and 3D FE simulation of
[30] Ares JR, Cuevas F. Mechanical milling and subsequent annealing effects on the Vickers indentation. J Alloy Comp 2018;752:137–47.
microstructural and hydrogenation properties of multisubstituted LaNi5 alloy. Acta [44] Atrian A, Majzoobi GH, Enayati MH, Bakhtiari H. Mechanical and microstructural
Mater 2005;53:2157–67. characterization of Al7075/SiC nanocomposites fabricated by dynamic
[31] Kang S, Jung YS, Jun JH, Lee YK. Effects of recrystallization annealing temperature compaction. Int J Miner Metall Mater 2014;21(3):295–303.
on carbide precipitation, microstructure, and mechanical properties in
Fe–18Mn–0.6 C–1.5 Al TWIP steel. Mater. Sci. Eng. A 2010;527(3):745–51.

Das könnte Ihnen auch gefallen