Você está na página 1de 15

Materials and Design 65 (2015) 11211135

Contents lists available at ScienceDirect

Materials and Design


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

Aluminium reinforced by WC and TiC nanoparticles (ex-situ)


and aluminide particles (in-situ): Microstructure, wear and corrosion
behaviour
A. Lekatou a,, A.E. Karantzalis a, A. Evangelou a, V. Gousia a, G. Kaptay b, Z. Gcsi b, P. Baumli b, A. Simon b
a
b

Department of Materials Science and Engineering, University of Ioannina, Ioannina 45110, Greece
Physical Metallurgy Department, Materials Science Institute, The University of Miskolc, H-3515 Miskolc-Egyetemvaros, Hungary

a r t i c l e

i n f o

Article history:
Received 11 June 2014
Accepted 14 August 2014
Available online 27 August 2014
Keywords:
Aluminium Matrix Composites
WC/TiC nanoparticles
Melt inoculation
Sliding wear
Cyclic polarization
Dilute Harrisons Solution

a b s t r a c t
In the present effort, Aluminium Matrix Composites (AMCs) were produced by the addition of submicron
sized TiC and WC particles of low (up to 1.0 vol%) content into a melt of Al1050. Casting was assisted by
the use of K2TiF6 as a wetting agent and mechanical stirring to limit particle clustering. An extensive
presence of intermetallic phases was observed in the cast products, as a result of both the inoculation
by K2TiF6 and the intensive mainly due to the ne carbide particle size reactivity of the carbides with
the molten matrix. Particle distribution was reasonably uniform comprising both clusters and isolated
particles. The intermetallic particle dispersion has changed the intended nature of the composites.
Instead of one type of reinforcement, that of carbide particles, the aluminium matrix contained two main
types of reinforcement: (a) in-situ intermetallic particles and (b) carbide nanoparticles, as such, or more
often as clusters of remaining carbide nanocores and aluminide particles. The reinforced materials
exhibited a notably improved sliding wear performance over that of the alloy owing to the benecial
effect of both the carbide and the intermetallic phase dispersion. A wear mechanism was formulated
based on microstructural features of the wear surface (repeated hill-valley morphology, surface oxide
layers, crack formation and grooving). Cyclic potentiodynamic polarization in Dilute Harrisons Solution
(DHS) revealed that the corrosion behaviour of the reinforced materials was mainly controlled by the
corrosion of the alloy matrix. As such, the predominating form of corrosion was intergranular corrosion
(IC) of Al associated with the presence of alloy matrix impurities. Carbide nanoparticles, aluminide phase
associated with them and their Al-matrix remained essentially intact of corrosion. IC progress was often
inhibited by the presence of clusters of aluminide and carbide particles.
2014 Elsevier Ltd. All rights reserved.

1. Introduction
Due to properties, such as high specic strength and stiffness,
low density and low thermal expansion coefcient, Aluminium
Matrix Composites (AMCs) have attracted great scientic attention
as candidate materials for high-tech, structural and functional
applications including aerospace, defense, automotive, electronic
packaging, precision instruments, thermal management areas,
sports equipment and recreation. Among them, particulate
reinforced AMCs (PRAMs) have extensively been investigated
owing to their low production costs and versatility in employing
conventional techniques for their production and shaping [13].
Recent evidence has shown that the mechanical response of AMCs
can further be improved if submicron- and nano- sized particles
Corresponding author. Tel.: +30 26510 07309; fax: +30 26510 07034.
E-mail address: alekatou@cc.uoi.gr (A. Lekatou).
http://dx.doi.org/10.1016/j.matdes.2014.08.040
0261-3069/ 2014 Elsevier Ltd. All rights reserved.

are used as reinforcing phase [47]. Within the above framework,


the main concept behind this work is to exploit the attractive
properties of aluminium in nanotechnology by fabricating AMCs
reinforced by submicron sized carbide particles and evaluating
them in terms of their corrosion and wear behaviour. However,
there are some major drawbacks for such concept applicability:
(a) the high production costs involved in several production routes,
(b) the agglomeration of nanoparticles when adopting low cost
casting techniques and (c) the likely degradation of corrosion
resistance due to the introduction of a high number of interfaces.
Regarding the rst drawback, amongst the great variety of
manufacturing methods adopted for AMCs, the conventional
casting processes are always at the research forefront due to their
relatively low cost, ease-to-handle advantages and large scale
production capabilities. Works on cast nano-particle reinforced
AMCs have mainly involved Al2O3 [6,812], SiC [7,13], B4C [14],
AlN [15], MgO [16].

1122

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

Regarding the second drawback, in conventional casting processes, the particle-molten Al wetting behaviour is the most crucial
factor for a successful particle insertion in the melt. Towards this
direction, ceramic phases with a strong metallic character, such
as TiC and WC, may ensure enhanced particle liquid metal wetting compatibility [17]. Besides, TiC and WC exhibit very high
hardness and modulus of elasticity, excellent wear and high temperature properties and good corrosion resistance, properties of
great industrial signicance. Nevertheless, research efforts on WC
and TiC cast reinforced AMCs are limited and they, in their vast
majority, concern microsized TiCp reinforced AMCs. These works
have produced promising results as far as the wetting behaviour
of the melt, the dispersion uniformity, as well as the strength
and wear resistance are concerned [1825]. Additionally, the particle insertion can further be assisted by the use of uxing agents,
such as halide salts that dissolve the oxide layer formed on the
surface of the molten alloy; thus, the involved phases are allowed
to express their net wetting characteristics and enhance particle
incorporation [21,22,24,2629].
Regarding the third drawback, it is well established that the
aqueous corrosion behaviour of AMCs can be affected by many
AMC features. The matrix-reinforcement interface, the matrix/
secondary phase interface, the secondary phase nature and content,
the reinforcement/secondary phase interface, the electrolyte active
ion concentration, the production route, the heat treatment, the surface treatment can signicantly affect the corrosion resistance and
mechanisms [3034]. Owing to these too many factors, conicting
data and interpretations exist regarding fundamental issues, such
as corrosion resistance and corrosion initiation sites [35]. Several
studies have reported lower corrosion resistances for AMCs in comparison with the respective monolithic alloys. Various reasons have
been considered responsible for this deterioration of corrosion resistance in AMCs: (a) breakdown of a continuous passive lm at the
matrix/reinforcement interfaces [36]; (b) galvanic coupling of aluminium and reinforcement [37]; (c) voids at the reinforcement/
matrix interface [38]; (d) an increase in the dislocation
density around particle clusters [39]; (e) interfacial layers around
particulate reinforcements that promote galvanic corrosion [30];
(f) formation of intermetallic phases by reaction of the reinforcement with the matrix due to heat treatment [34] or precipitation
of intermetallic phases enhanced by the presence of particulates
[33]. More recently, Pardo et al. [40] noted that the corrosion damage in AA360/SiCp and AA380/SiCp composites in (13.5) wt% NaCl
was caused by pitting attack mainly at the reinforcement/matrix
interface.
However, Grifths and Turnbull [41] did not notice any apparent effect of SiC reinforcement on the electrochemical behaviour
of Al6061 in aerated 3.5 wt% NaCl. They concluded that the effects
of reinforcement on the corrosion of Al cannot be generalized and
are specic functions of the environmental conditions and the processing route. The following works tend to support this claim:
Trowsdale et al. [38], whilst nding no signicant galvanic action
between SiC and Al, noticed that 20 wt% incorporation of SiC (of
particle size of 3 lm) in Al1050 led to a slight reduction of the pitting resistance of the alloy. However, large particle sizes (20 lm)
led to intensication of pitting due to cracking during the fabrication process. Kiourtsidis et al. [42] stated that the SiCp presence
does not accelerate failure of the passivation oxide lm, whereas
pitting corrosion potentials in aerated 3.5 wt% NaCl are not
considerably affected by the SiCp content at a given aging condition; nevertheless, alteration in aging kinetics due to SiCp presence
is responsible for the differentiation in the pitting corrosion
behaviour among the composites. Alaneme and Bodunrin [43]
claimed that unreinforced AA6063 exhibited slightly superior corrosion resistance than the AA6063/Al2O3p composites in NaCl and
NaOH media; however, the composites showed better corrosion

resistance in H2SO4 medium. Candan [44] noted that intermetallics


as a result of reaction between an AlMg alloy and SiC reinforcement beneted the corrosion resistance of the composites in 3.5%
NaCl due to interruption of the continuity of matrix channels.
On the other hand, a number of works have reported superior
corrosion performance of the particulate AMCs in relation to the
respective alloy. The corrosion resistance of LM13-Al to 1 M HCl
was found by Seah et al. [45] to improve with increasing the garnet
particle content (26 wt% garnet) due to the garnet particles acting
as physical barriers to the corrosion process. The positive effect of
zircon (ZrSiO4) particulates on the corrosion resistance of Al6061
(HCl of different concentrations) and Al7075 (seawater) has been
reported by Jameel et al. [46] and Nagaswarupa et al. [47], respectively. Toptan et al. [48] noted that (1519) vol% addition of B4C
particles did not signicantly affect the tendency for corrosion of
an AlSiCuMg alloy in 0.05 M NaCl; however, it decreased the
corrosion tendency and corrosion rate during sliding wear testing
in 0.05 M NaCl.
The wear performance of AMCs is of major importance, since it
has to satisfy the necessity for long-lasting applications. The sliding
wear behaviour of PRAMCs has extensively been investigated during the last decades [4955]. Key features of the involved degradation mechanisms include: (i) material parameters, such as matrix
microstructure and hardness, particle size and shape, particle nature, particle volume fraction, particle matrix interface integrity,
interfacial bonding, reinforcement wettability by the matrix, secondary phase particles; and (ii) extrinsic tribological parameters,
such as externally applied normal load, sliding speed, sliding distance, temperature, surface nish, hardness of counterpart and
nominal contact area [54,56]. Owing to these too many factors,
no consistent wear behaviour of AMCs has been established [52].
Surface oxidation, extensive plastic deformation, debris characteristics and nature can each play a crucial role that can vary the wear
mode form mild to severe [49,50,53,57]. The governing mechanisms have been described in three established theories: (a) the
adhesive wear theory that considers the adhesion between the
counter surfaces at the asperities and subsequent decohesion of
asperities leading to material removal [58]; (b) the delamination
wear theory taking place in four steps: cyclic plastic deformation
of surface layers, crack or void nucleation, crack growth, formation
of debris and debris removal by extension of cracks to the surface
[57]; (c) the mechanically mixed layer theory, involving formation
of debris owing to oxidation and plastic deformation of the counter
surfaces [54,59]; this debris is continuously in a state of comminution and consolidation and, eventually, forms a hard protective
surface layer that reduces the overall wear rate. On numerous
occasions, the above theories complement each other. According
to Al-Qutub [60], erosion wear dominates in mild wear conditions;
delamination wear becomes the primary mode in the wear transition state; the severe wear regime is governed by adhesion wear
(submicron Al2O3p/Al6061). Sub-micron and nano particulate
ceramic reinforcements, such as Al2O3 [60,61], B4C [62], MoSi2
[63], SiC [64], SiC/graphite [65] and TiC [66], have been found to
improve the wear resistance of aluminium (AMCs prepared by
powder metallurgy techniques).
Overall, the main objective of the present effort is to fabricate
PRAMCs by adopting four approaches: (a) low cost conventional
casting assisted by stirring and salt uxing for improved particle
wetting and distribution, (b) addition of submicron TiC and WC
particles as the primary reinforcement in order to combine the
advantages of ultra ne dimensions with the excellent intrinsic
properties of these carbides, (c) employment of low primary
reinforcement volume fractions to limit segregation, (d) optimization of surface property response by attaining further in-situ
reinforcement (whilst at the same time keeping production costs
low).

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

2. Experimental procedure
AMCs were prepared by the addition of sub-micron sized WC
particles (of approximate particle size of 200400 nm) and TiC
particles (of approximate particle size of 400700 nm) into
Al1050 (Al of 99.5% commercial purity). The compositions
employed were: Al-0.7 vol% TiC, Al-1.0 vol% TiC, Al-0.5 vol% WC
and Al-1.0 vol% WC. Wetting and homogenization were assisted
by two approaches: Fluxing and mechanical stirring. K2TiF6 (10 g
K2TiF6/190 g Al1050) was utilized as a uxing salt for removing
the oxide phase from the surface of the aluminium melt [26,29].
First, mixing of the reinforcement and the salt was carried out. Then,
this mixture was added into the alloy melt (830 C). The salt was
allowed to react with Al, a slag was formed, the carbide particles
inltrated into the melt and the slag was removed by a ladle. Rigorous stirring was then applied for homogenization and breakage of
any initial particle clusters. Stirring was conducted by an in-house
made apparatus based on an AEG SB2E 700R power drill, with a
graphite rod being adapted and a four branch stirring shaft being
assembled at the end of the rod. The stirring speed was kept
constant at 3200 rpm. The stirring duration was 20 s. A nal slag
cleaning was performed prior to casting into cylindrical, steel
moulds of 1.5 cm inner diameter and 15 cm height.
Specimens were cut from each cast bar, mounted and prepared for
metallographic examination. Standard metallographic procedures

Temperature (K)

1123

were carried out, which included grinding by SiC papers followed


by polishing with diamond suspensions. Inspection of all samples
was performed by Scanning Electron Microscopy (JEOL JSM 6510LV
SEM/ Oxford Instruments X- Act EDX).
Specimens polished to Ra < 1 lm were subjected to dry sliding
wear testing at room temperature. A ball-on-disk tribometer (CSM
Instr.) was employed. The following parameters were employed:
normal load of 1 N, sliding speed of 10 cm/s, acquisition rate of
20 Hz, total sliding distance of 1000 m. AISI 5210 steel balls of
6.0 mm diameter were used as a counterbody material. Each run
was interrupted every 200 m, for measuring the mass loss of the
sample (Toledo electronic balance of ve decimal digits). Before
each weighing, the specimen was ultrasonically cleaned by acetone.
The overall wear rate was calculated from the mass loss vs. sliding
distance data by linear regression analysis (least squares method).
Triplicate tests were performed for each material type.
Corrosion testing was conducted on cylindrical coupons that
were cut with a diamond saw, ground to 1000 grit, ultrasonically
cleaned and encapsulated in PTFE, leaving a surface area of
1 cm2 to be exposed to aerated DHS, at 25 C. DHS (Dilute Harrisons Solution) is a testing solution often used on aeronautical
alloys to approach atmospheric conditions often encountered by
airplanes [67]. The solution contains ammonium sulphate and
sodium chloride (0.35 wt% (NH4)2SO4 + 0.05 wt% NaCl) usually
found in atmospherically deposited moisture. DHS is an effective
emulator of the effects of acid rain. All the electrochemical tests
were performed using the Gill AC potentiostat/galvanostat by
ACM Instruments. A standard three electrode cell was employed,
with Ag/AgCl (3.5 M KCl) as the reference electrode and a platinum
gauge as the counter electrode. Potentiodynamic polarization tests
were carried out at a scan rate of 10 mV/min. Polarization scanning
started after 4 h of recording the Open Circuit Potential in DHS, at
25 C. Reverse polarization was conducted to study the susceptibility of the specimens to localized corrosion.

3. Results and discussion


3.1. Composite microstructure particle incorporation

Al

Ti

Al (mole fracon)

First of all, the interpretation of the results herein presented,


utilizes the AlTi [68], AlW [69], TiW [70] and AlC [71] phase
diagrams. The AlTi and AlW phase diagrams are illustrated in
Fig. 1.
Fig. 2 presents the microstructures of the TiCp/Al materials
(Fig. 2a and b) and WCp/Al materials (Fig. 2c and d). It is seen, that
the carbide phase is present as both isolated particles and particle
clusters, which are mainly located at the grain boundaries. In all
cases, there is a strong presence of intermetallic phases, located
in the interior and at the boundaries of the Al grains.

Temperature (K)

liquid

Al2W

bcc

Al7W3

Al

Al4W

fcc

Al12W
Al5W

Al77W23

W (mole fracon)

Fig. 1. Phase diagrams of the systems: (a) TiAl [68], e(l) = Al3Ti; and (b) AlW [69].

3.1.1. TiCp reinforced alloy


In the case of the TiCp reinforced alloy, the intermetallic phase
present has been identied by quantitative EDX as Al3Ti, which is
in consistency with the Al-Ti phase diagram (Fig. 1a). Fig. 3 shows
that the Al3Ti phase presents three main morphologies: (a) blocky,
longish rectangular plates/rods (e.g. particles from which spectrum
10 in Fig. 3a, spectrum 19 and spectrum 23 in Fig. 3b have been
received); (b) large, rounded particles of diameter (16) lm (e.g.
particles from which spectra 8, 9, 11 in Fig. 3a, spectra 17, 18
and 20 in Fig. 3b have been received); (c) dispersions of ne
rounded or plate-like precipitates (Fig. 3a and c). (Typical EDX
analyses are included in the legend of Fig. 3). The ne Al3Ti
precipitates are usually associated with the presence of the TiC
nanoparticles or their remaining cores (circled cluster in Fig. 3c).

1124

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

0.7 vol% TiC

0.5 vol% WC

1.0 vol% TiC

1.0 vol% WC

Fig. 2. Microstructures of the cast composites (backscattered electron-BSE mode). Isolated and clustered carbide particles are noted by arrows and circles/ellipses,
respectively.

0.7 vol% TiC

1.0 vol% TiC

Al3Ti

Al3Ti

Al3Ti
Al3Ti

1.0 vol% TiC

Fig. 3. Intermetallic compound particles observed in the TiCp/Al materials (a & b: Secondary Electron-SE mode, c: BSE mode). Spot EDX analyses in (at%): (a) spectrum 8:
76.97 Al-23.03 Ti, spectrum 9: 76.89 Al-23.11 Ti, spectrum 10: 75.97 Al-24.03 Ti, spectrum 11: 76.56 Al-23.44 Ti; (b) spectrum 18: 76.16 Al-23.84 Ti, spectrum 19: 76.06 Al23.94 Ti, spectrum 20: 76.68 Al-23.32 Ti; (c) Al3Ti formation associated with TiC nanoparticles (in circle).

The origin of the Al3Ti formation can be twofold: (a) due to reaction between K2TiF6 and molten Al [72,73] and (b) due to reaction
between molten Al and TiC. Taking into account the TiAl (Fig. 1a)
and AlC phase diagrams, the following reactions may account for
the presence of Al3Ti particles in the TiCp reinforced materials:
At the temperature of the melt (830 C):

3K2 TiF6 s 13All ! 3Al3 Tis 3KAlF4 l K3 AlF6 l 72

13All 3TiCs ! 3Al3 Tis Al4 C3 s

Upon cooling, at the Al/TiC interface, the remaining Al(l) reacted


with Al3Ti by a peritectic reaction:

All Al3 Tis ! Alsperitectic; 666  C

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

The peritectic mode of Al3Ti engulfment by the growing aAl grain


may account for the frequent location of Al3Ti inside the Al grains,
as observed in Fig. 3.
Reaction (1) is expected, as the outcome of a standard Al inoculation process described by Mahallawy et al. [72]. Reaction (2) has also
been reported to occur at temperature levels that include the melt
temperature of the present effort [74] or even at lower
temperatures upon cooling, despite the fact that this reaction is thermodynamically unlikely [74]. However, previous works with AMCs
reinforced by TiCp of conventional particle size (45 lm) at the
same processing temperature did not show any interaction between
Al and TiC [18,21,25]. Therefore, it is inferred that the submicron size
of the TiC particles has accelerated reactivity phenomena owing to
the high number/specic surface of the interfaces introduced.
The non-detection of Al4C3 in the nal products can be
explained by its dissolution and removal during metallographic
preparation, since Al4C3 is hydrolyzed by water according to the
reaction [75]:

Al4 C3 s 12H2 Ol 4AlOH3 3CH4 g

The different Al3Ti particle morphologies can be explained on the


basis of reactions (1)(3): Blocky, longish plates are most likely
the product of the reaction between K2TiF6 and Al(l) at the temperature of the melt (830 C). These morphologies are usually formed
in salt rich regions, i.e. Ti-supersaturated regions [72]. Large
rounded particles can also be the product of salt-melt reactions in
areas of somewhat lower K2TiF6 concentration, since the morphology of Al3Ti particles resulting from the salt-Al(l) reaction is
strongly dependent on both the Ti content in the melt and the
cooling rate; thus, a wide range of different morphologies may be
attained [72]. The ne precipitates associated with TiC nanoparticles are probably the product of reaction (2).
3.1.2. WCp reinforced alloy
In the case of WCp/Al, two types of tungsten aluminide are
observed (Fig. 4): Al12W in the form of coarse polygonal particles
of largest diagonal of (318) lm (Fig. 4a) and Al5W in the form
of acicular plates (Fig. 4b and c). (Their stoichiometry has been
identied by quantitative EDX. Typical EDX analyses are included
in the legend of Fig. 4). Their presence is fully justied by the
AlW (Fig. 1b) and AlC phase diagrams, as follows: at the temperature of the melt (830 C), Al(l) combined with WC to form Al5W
and Al4C3. (Al5W needles are always located by WC remaining
cores, as shown in Fig. 4b and c). On cooling, Al12W was formed
as a result of the peritectic reaction between Al(l) and Al5W. On
further cooling, the remaining Al(l) peritectically reacted with
Al12W to form aAl. The peritectic mode of Al12W engulfment by
the growing aAl grain may account for the localization of Al12W
in the interior of the Al grains. Therefore, the following sequence
of reactions may account for the presence of Al5W and Al12W particles in the WCp reinforced materials.
At the melt temperature830  C : 19All 3WCs ! 3Al5 Ws Al4 C3 s
5

On cooling697  C : All Al5 Ws ! Al12 Wsperitectic

On cooling661  C : All Al12 Ws ! Alsperitectic

Al3Ti particles appear as clusters or agglomerates of ne rounded


particles or as a system of coarse rectangular plates forming an
incomplete rosette (Fig. 4c, upper right). As aforementioned in
Section 3.1.1, a variety of different aluminide morphologies can
result from the interaction of the Al-melt with the salt ux. The
morphology of incomplete rosette may be associated with salt rich
regions, i.e. Ti-supersaturated regions [72]. The signicant presence
of W in the Al3Ti composition (see spectrum 13 in Fig. 4a, spectra 2,

1125

3, in Fig. 4c) can be explained by the fact that, at the temperature of


the melt (830 C), Ti and W can form (Ti,W) solid solutions (b1, b2, b)
over a wide range of stoichiometry [70]. During casting and
solidication, a fair amount of W has remained trapped in the Al3Ti
lattice resulting in a metastable Al3(Ti,W) structure. It is known that
Al3Ti of tetragonal D022 structure can be transformed to the
metastable L12 cubic structure by alloying with transition metals
with more d-electrons in their valence band (e.g. W). The L12
structure is more ductile than the D022 structure due to covalent
bonding with enhanced metallinity as compared to the covalent
bonding of the D022 structure [76].
The great reactivity of WC with molten Al is not only due to its
submicron particle size (ner than that of TiCp) but also due to its
thermodynamic instability. {The enthalpy of formation of WC is
higher than that of TiC {DHf(WC): 40.5 kJ/mol, DHf(TiC):
184.1 kJ/mol [77])}.
3.1.3. Al-Fe based intermetallics
Another important feature, observed in Fig. 2 and, in a high
magnication, in Fig 5, is the presence of eutectic phase at the
grain boundaries. EDX analysis revealed that the eutectic microconstituent, apart from Al, consists of AlFe and AlFeSi intermetallic phases. Such eutectic presence is commonly encountered in
commercial Al-alloys [78]. Fe is the most common impurity in
aluminium forming a variety of intermetallics, such as Al3Fe, Al6Fe,
aAl(Fe, Mn)Si, dAlFeSi, b(Fe,Si), a(Fe,Si), Al12Fe3Si2.
3.1.4. Particle distribution
Despite the reaction product extent in both types of composite
materials, it could be stated that the particle incorporation within
the melt was successful and of high rate. Such a successful incorporation is attributed to both: (a) the benecial action of K2TiF6,
which reacted with liquid Al to form a KAlF based liquid slag
that removed surface oxide phases and allowed the expression of
the particle-melt net wetting characteristics; and (b) the stirring
applied during processing.
The carbide nanoparticle distribution, characterized -as aforementioned- by clustered and isolated particles located mainly in
the vicinity of the grain boundaries, could be a result of: (a) initial
particle clustering of the precursor powder that could not be broken by mechanical stirring. The very ne particle size enhances
such clustering endurance; (b) cluster formation due to particle
pushing by the solidication front. Such pushing could also be
responsible for the nal particle location at the grain boundaries
and the areas of the lastly solidied liquid. Thermal conductivity
difference theories, proposed by Zubko et al. [79] and Surappa
and Rohatgi [3], sufciently describe this nal particle location.
According to these theories, during cooling, the hotter due to
their lower thermal conductivity reinforcing particles preserve
the cooler surrounding liquid. This way, the growth of an advancing grain upon cooling, is obstructed by the reinforcing particles. As
a consequence, aluminium grains become rened and particles are
being pushed towards the grain boundaries. Thermal conductivity
theories can also explain the Al3Ti localization at grain boundaries.
On the other hand, the localization of Al3Ti inside aAl grains can be
explained by the peritectic mode of Al3Ti engulfment by aAl
according to reaction (3). In the WCp reinforced alloy, the presence
of Al3Ti at grain boundaries (Fig. 4a and c) is more frequent than in
the TiCp reinforced alloy, because less Al(s) was available to peritectically engulf Al3Ti; Al(l) -reactant in reaction (3)- had largely
been consumed in reactions (5)(7).
Here it should be noted that the term carbide nanoparticle actually refers to the cores of the original sub-micron carbide particles
remaining after the reaction of their peripheries with the Al-melt.
To conclude, the distribution of intermetallic particles, isolated
carbide nanoparticles and clusters of intermetallic/nanocarbide

1126

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

0.5 vol% WC

0.5 vol% WC

Al3Ti

Al5W
Al3Ti

WC

Al3Ti

Al3Ti
Al12W

1.0 vol % WC

Al3Ti
Al5W

Al3Ti
WC

Al5W
Fig. 4. Intermetallic compound particles observed in the WCp/Al materials (SE mode). Spot EDX analyses in (at%): (a) spectrum 12: 91.73 Al-7.64 W-0.63 Ti, spectrum 13:
76.01 Al-18.47 Ti-5.52 W, (b) spectrum 29: 83.88 Al-15.55 W-0.57 Ti, and (c) spectrum 2: 75.44 Al-19.05 Ti-5.51 W, spectrum 3: 75.23 Al-18.95 Ti-5.82 W, spectrum 5: 83.86
Al-15.03 W-1.11 Ti.

1.0 vol% TiC

Fig. 5. Intergranular presence of a eutectic (iron aluminide) intermetallic phase (SE


mode).

particles was reasonable uniform. However, it changed the


intended nature of the composites. Instead of one type of reinforcement, that of carbide particles, the AMCs contained two main types
of reinforcement: (a) in-situ Al3Ti or (Al5W + Al12W + Al3(Ti,W))
intermetallic particles in the cases of TiCp-AMC and WCp-AMC,
respectively; (b) carbide nanoparticles, as such, or more often as
clusters of carbide nanoparticles and aluminide particles. A main
question is arising: Had the new composite improved surface properties in relation to the monolithic alloy?
3.2. Sliding wear response
3.2.1. Effect of carbide volume fraction and type
Fig. 6a, presents the mass loss of the different composites as a
function of the sliding distance. It is seen that, as the sliding

distance increased, the mass loss also increased, in compatibility


with previous investigations [51,52,56,80,81]. The wear rates of
the produced materials are displayed in Fig. 6b. It is evident that
increasing the carbide particle volume fraction has led to a
decrease in mass loss, and, consequently, wear rate, which is in
agreement with other experimental efforts [18,49,50,52,53,80,82].
The positive effect of the carbide volume fraction on the wear
resistance of the composites is both direct and indirect: the direct
effect stems from the TiC and WC particles, as such and as clusters;
the indirect effects originates from the hard intermetallic compound particles formed by the reaction of the carbide phase with
the Al-matrix. This dual effect has led to a notable decrease in
the wear rate with carbide volume fraction increasing despite the
low carbide volume fractions employed. In particular, additions
of just 0.5 vol% WC and 1.0 vol% WC have led to a decrease in the
wear rate of Al1050 by a factor of 2.7 and 3.7, respectively; additions of 1.0 vol% TiC have led to a decrease in the wear rate of
Al1050 by a factor of 2.2. Such benecial behaviour is mainly
attributed to the strengthening effect that the dispersed particles
(carbides and aluminides) induce to the soft matrix, delaying, in
turn, plastic deformation phenomena which can be mainly responsible for the overall degradation sequence. More analytically,
carbide particles (in their majority as remaining cores of nanodimensions) and intermetallic particles may inhibit/retard plastic
deformation-due crack growth in the matrix by: (a) reducing the load
transfer to the matrix, (b) decreasing the direct matrix-counterface
contact area, (c) providing thermal stability to the matrix, thus
postponing thermal softening effects and (d) inducing Al-grain
renement (see Section 3.1.4) [18,49,50,5456,80]. Regarding contribution (a), the particle clusters (nanocarbide or nanocarbide/
intermetallic) and the coarse intermetallic particles (e.g. Al12W)
may carry great portions of the applied load, thereby reducing
the load that ne particles and the soft matrix can carry [49]. On
the other hand, the probability of reinforcement cracking increases
with increasing size when size exceeds a critical value [83].

A. Lekatou et al. / Materials and Design 65 (2015) 11211135


40
35
-3

Mass loss (g x10 )

Al-1.0 vol% WC
Al-0.5 vol% WC

30

Al-1.0 vol% TiC

25

Al1050

Al-0.7 vol% TiC

20
15
10
5
0

200

400

600

800

1000

1200

Sliding distance (m)

Fig. 6. (a) Mass loss versus sliding distance during dry ball-on-disk testing, for Al
1050, monolithic and reinforced by TiC and WC submicron particles; (b) the wear
rate of the different materials.

Fig. 6 also demonstrates that WC reinforcement has led to a


higher wear resistance of the WCp reinforced alloy as compared
to the TiCp reinforced alloy. This superiority can be attributed to:
(a) the ner particle size of WC resulting in a greater obstruction
of the dislocation movement due to the greater number of
interphase boundaries, and (b) further densication of the hard
phase dispersion by the more extensive presence of intermetallic
compound particles (Al5W, Al12W); the latter, as aforementioned
in Section 3.1.2, is due to the higher reactivity of WC as compared
to that of TiC (ner particle size and lower thermodynamic stability in comparison with TiC).
3.2.2. Wear track morphology
The wear track topographies of the monolithic alloy and the
AMCs are illustrated in Fig. 7. In all cases, the characteristic hillvalley morphology is observed. The wear track appears wider at
the hill areas and narrower at the valleys; this indicates that,
during sliding, signicant material movement towards the hills
has occurred. Such landscape formation has been explained by
Sarkar [84] in terms of intensive plastic deformation of the soft
matrix in front of the moving counterbody steel ball causing
notable material ow at directions, which are both parallel and
perpendicular to the sliding direction. As a result, a hill is being
built up. When the counterface movement cannot any longer cause
further material ow to the hill, the counterface overpasses it
and repeats the same cycle on an adjacent area. Eventually, a
new hill is being formed. The nal outcome is the repeated
hill-valley morphology observed in Fig. 7.
Comparison of the wear track morphologies of the reinforced
materials with that of the monolithic alloy in Fig. 7 draws the
following observations: (a) the hill-valley morphology of the
monolithic alloy is more intensive than that of the reinforced

1127

materials, as far as the hill width to valley width ratio and


the wear track relief are concerned; (b) the prole of the wear track
edges of the monolithic alloy is rougher than those of the AMCs; (c)
the unreinforced alloy shows the widest wear track, whereas the
1 vol% WC composite shows the narrowest wear track. Comparison
of the wear track morphologies of aluminium reinforced by the
same type of carbide in Fig. 7 reveals that the higher the carbide
volume fraction the more uniform the landscape morphology and
the less rough the wear track edges. It is, thus, evident that the
monolithic alloy has been subjected to more severe plastic
deformation than the reinforced alloy. Therefore, it is deduced that
the reinforcing phases (carbides and aluminides) have restricted
the matrix plastic ow resulting in a more uniform wear track
landscape [18].
EDX analysis on selected hill areas, illustrated in Fig. 8a and b,
revealed the presence of Al-based oxide phases with Ti (TiCp/Al)
and W (WCp/Al) also being present. Representative EDX spectra
are given in Fig. 8c and d. The formation of oxide layers during sliding wear of Al-alloys and their composites has previously been
reported [18,50,53,85]. Their presence is owing to a mechanical
mixing process accompanied by oxidation reaction due to the frictional heating during dry sliding. Formation of these oxide layers
delays the transition between mild and severe wear regimes in
AMCs.
Higher magnication micrographs of the wear surfaces, in
Fig. 9, demonstrate a quite greater extent of plastic deformation
for Al1050 (Fig. 9a), in comparison with the TiC and WC reinforced
alloy (Fig. 9b and c, respectively). In fact, the wear surface of the
monolithic alloy exhibits mainly plastic deformation (in terms of
hills/valleys, ridges and, generally, surface relief). The wear surfaces of the composites show both the above type of plastic deformation, as well as groove formation (aligned in the sliding
direction) with the latter being the main feature. The grooves on
the wear surfaces of the AMCs are shallow, ne and narrowly
spaced, a typical pattern caused by the abrasive action of large
numbers of hard particles and debris [50,81]. The 0.5 vol% WC
composite presents slightly more extensive grooving as compared
to the 1.0 vol% TiC composite indicating a sliding action by a higher
number of hard particles (despite a lower volume fraction). The
above observation will be discussed in Section 3.2.3.
High magnication micrographs of the wear surfaces, in Fig. 10,
reveal the presence of cracks/aws within the oxide layer almost
perpendicular to the sliding direction. Crack formation during wear
testing of AMCs has been reported to be caused by extensive plastic deformation of the Al-matrix and/or particlematrix interface
debonding, eventually leading to delamination and material
removal [20,54,56,57]. Another likely reason for such crack formation may be -besides fatigue due to repeated sliding- thermal fatigue of the oxide phases due to thermal cycling, as the test was
interrupted every 200 m of sliding distance for specimen weighing
[18,82]. Cracking in the surface oxide layer and subsequent delamination constitute crucial events for the wear response of an Albased material, since they have been considered responsible for
the transition between mild wear and severe wear regimes
[81,86]. Comparison of Figs. 10ac shows that the monolithic
matrix exhibits the biggest cracks that may lead to delamination
(as the one observed at the far left of Fig. 10a). Comparison of
Fig. 10b and c reveals a smoother wear surface, fewer cracks/surface damage and ner, shallower grooves for the 0.5 vol% WC composite in relation to the 0.7 vol% TiC composite.
Overall, the wear mode of the produced AMCs is characterized
as mild, as it lacks any signicant seizure, material delamination
or extensive material deformation. The mild wear regime is attributed to the benecial action of two main factors: (a) the extensive
hard phase (carbides and intermetallics) dispersion has restricted
matrix ow and delayed crack growth and (b) the surface oxide

1128

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

monolithic

valley
hill

0.7 vol% TiC

1.0 vol% TiC

valley
hill

hill

valley

0.5 vol% WC

1.0 vol% WC

hill
valley

valley
hill

Fig. 7. Panoramic views of the wear track morphologies of the monolithic alloy and the different composites produced (SE mode), illustrating the hill-valley landscape.

1.0 vol% TiC

1.0 vol% WC

Al

Al

O
Ti

Fig. 8. Wear track morphologies in BSE mode and EDX spectra from hill (dark contrast) areas. (a) and (c) Al-1.0 vol% TiC and EDX spectrum from a hill, respectively; (b) and
(d) Al-1.0 vol% WC and EDX spectrum from a hill, respectively.

1129

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

monolithic

1.0 vol% TiC

0.5 vol% WC

Fig. 9. Wear track morphologies of different materials produced (SE mode), illustrating a high extent of plastic deformation for the monolithic alloy and the formation of ne
and dense grooves on the wear surface of the AMCs.

monolithic

0.5 vol% WC

0.7 vol% TiC

Al

O
W

Fig. 10. (a)(c): Crack formation in oxide layer areas (SE mode); (d) EDX analysis in the vicinity of the cracks revealing Al-based oxide presence (Al-0.5 vol% WC).

layer has protected underlying matrix areas from direct contact


with the counterbody material.
3.2.3. Wear debris
Comparison of the micrographs of the WC reinforced alloy with
those of the TiC reinforced alloy and the monolithic alloy, shows
that the wear tracks and the neighbouring external zones of the
WC-AMCs present notably more debris than the corresponding
surface zones of the TiC-AMCs and the monolithic alloy (compare
Figs. 7d and e with Figs. 7ac, Fig. 8b with Fig. 8a, Fig. 9c with

Figs. 9a and b, Fig. 10c with Figs. 10a and b). This observation is
compatible with the relatively extensive (ne and dense) grooving
along the wear track of the WC-reinforced alloy (compare Fig. 9c
with Figs. 9b and a).
EDX analysis of debris particles from the wear surface of the
WC reinforced alloy (Fig. 11a) revealed two types of materials:
(a) bright contrast particles of high W content (e.g. spectra 13
and 15) and darker contrast oxide particles of relatively low W
content (spectra 14, 16). It is indicated that the former ones
originated from oxidized broken particles of W-aluminides (most

1130

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

Al

1.0 vol% WC

Spectrum 13

W
W

Al

Spectrum 16

Ti

1.0 vol% TiC


Al

Spectrum 8

Ti

Fig. 11. Debris particles on the wear surface of the AMCs. (a) 1.0 vol% WC reinforced alloy and EDX spectra of brighter contrast particles (Spextrum13) and darker contrast
particles (Spectrum16); and (b) 1.0 vol% TiC reinforced alloy and EDX analysis from a debris particle (spectrum 8).

likely Al12W), whilst the latter ones originated from oxidized


mixtures of Al-matrix and fragments of W-, Ti- aluminides.
Al12W particles are highly susceptible to fragmentation not only
because of their large size but also because of their intrinsic brittleness; the latter is due to an ordered ve-fold icosahedral
structure, which is commonly met in quasicrystals and bulk
metallic glasses [87]. The acicular intermetallic compound particles (Al5W) also present a high tendency to fracture, especially
the ones lying normal to the surface; as such, they are likely
to participate in the relatively dark contrast debris particles
shown in Fig. 11a.
EDX analysis of debris from the wear surface of the TiC reinforced alloy (Fig. 11b) revealed only one type of material that of
alumina without any or with little Ti (spectrum 8); thus, it is indicated that debris from these composites mostly derived from the
Al-matrix mixed to a low extent with Al3Ti.
Based on the EDX analysis of debris, as aforementioned, and the
microstructural examination, as presented in Sections 3.1.1 and
3.1.2, the higher amount of wear debris in the case of the
WC-AMCs as compared to the TiC-AMCs, may be explained by
the more extensive presence of intermetallic particles that are
highly susceptible to fragmentation and can cause third body abrasion either as monolithic fragments or as mixtures with the alloy
matrix. In the case of WC-AMCs, third body abrasion has been conducted by Al12W (monolithic fragments with oxidized surfaces)
and acicular Al5W (mixed with alumina from the matrix and
Al3(Ti,W)).
Although the number of abrasive particles produced during
sliding wear is higher in the case of the WC-AMCs than in the case

of the TiC-AMCs, the relatively shallow and ne grooves along the


wear tracks in the case of WC-AMCs (compare Fig. 10b with
Fig. 10c) are evident of the relatively high resistance of WC-AMCs
to third body abrasion.
3.2.4. Mechanism of wear
Based on the aforementioned observations, a likely wear mechanism for the composites of the present effort can be summarized
as follows:
(i) On the onset of sliding wear testing, the soft aluminium
matrix was subjected to intensive plastic deformation in
front of the counterface ball. As a consequence, signicant
material ow occurred that, as sliding progressed, led to
the repeated hill-valley morphology. Carbide and intermetallic compound particles restricted the matrix ow resulting in smoother wear surfaces.
(ii) At the same time, the frictional heating during dry sliding
wear induced the formation of surface Al-based oxide layers,
enriched by Ti- or W- oxides.
(iii) The frictional forces, the brittleness of the alumina-based
surface layer and wear fatigue enhanced by thermal fatigue
(due to the repeated interruption of the test for mass loss
measurements) caused the formation of cracks in the oxide
layer.
(iv) Crack propagation and growth eventually led to material
removal. Crack propagation was delayed by the carbide/
intermetallic phase dispersion and the rened aAl grain
boundaries.

1131

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

3.3. Corrosion behaviour

1500

Potenal (mV, Ag/AgCl)

(v) The abrasive action of debris (TiCp/Al: oxides mostly originating from the matrix; WCp/Al: oxides originating from
mixtures of Al matrix and W-, Ti- aluminides and oxides
originating from fragments of large Al12W particles) caused
the formation of shallow, ne and dense grooves along the
wear tracks.

1000
500
0
-500
-1000
-1500
-2000

3.3.2. Microstructure of corrosion and correlation with polarization


behaviour
SEM micrographs of the corroded materials (cross sections) are
given in Fig. 14(ad). The main degradation morphology is intergranular corrosion (IC) associated with the presence of the AlFe/
Al or AlFeSi/Al eutectic microconstituent. In order to clarify the
predominant mechanism of corrosion, one has to initially consider
the generally accepted four steps involved in the localized corrosion of aluminium [88]:
(1) The adsorption of the reactive anions on localized sites of the
surface lm of aluminium (i.e. sites where the lm presents
inhomogeneities [89]). In the presence of incoherent precipitates, such as Al3Fe [90] and aAlFe(Mn)Si [91], the precipitate/Al interfaces would be the preferred sites for adsorption.
(2) The chemical reaction of the adsorbed anion with the aluminium ion in the aluminium oxide/hydroxide lattice.
(3) The thinning of the oxide lm by dissolution. This dissolution is a aw assisted/ aw centered process. (The passive
lm on Al-alloys exhibits semi-conductive properties owing
to the non-stoichiometry of composition and local structural
inhomogeneities [89]).
(4) The direct attack of the exposed metal by the anion possibly
assisted by an anodic potential.

0.00001

0.0001

0.001

0.01

0.1

10

100

10

100

10

100

10

100

Current density (mA/cm 2)


Potenal (mV, Ag/AgCl)

1500
1000
500
0
-500
-1000
-1500
-2000

0.00001

0.0001

0.001

0.01

0.1

Current density (mA/cm 2)

Potenal (mV, Ag/AgCl)

1500
1000
500
0
-500
-1000
-1500
-2000
0.00001

0.0001

0.001

0.01

0.1

Current density (mA/cm 2)


1500

Potenal (mV, Ag/AgCl)

3.3.1. Potentiodynamic polarization


The cyclic voltammograms of the tested materials are presented
in Fig. 12(ad). The negative hysteresis loops of the anodic polarization curves (i.e. higher current densities upon reverse polarization
as compared to the forward polarization) suggest that all materials
(monolithic alloy and AMCs) have been subjected to localized corrosion processes. The great similarity of the polarization curve shapes,
the nearly same areas of the hysteresis loops, as well as the similar
corrosion potential (Ecor) values and anodic-to-cathodic transition
potential (Ea/c tr) values (despite the different volume fractions and
particle reinforcements), indicate that the corrosion behaviour of
the tested materials was mainly controlled by the corrosion of the
monolithic alloy.
The only differences that are worthwhile to mention concern the
cathodic current density values recorded. These differences are
more clearly seen in Fig. 13, which includes only the cathodic part
of selected forward polarization curves. The WC reinforced materials present higher cathodic current densities compared with the
TiC reinforced materials (Figs. 12a and d and 13a). This can be associated with (a) the coarse Al12W particles, which have large enough
surfaces to sustain cathodic reactions, and (b) the increased area
fraction of noble intermetallic particles (Al12W, Al5W and Al3(Ti,W))
due to the relatively high reactivity of WC particles, as aforementioned in Sections 3.1.2 and 3.2.1. Furthermore, the alloy reinforced
with the high volume fraction of carbide phase presents higher
cathodic current densities than the alloy reinforced with the low
volume fraction of carbide phase, for the same carbide type
(Figs. 12b and c and 13b). This trend may also be associated with
the increased area fraction of noble intermetallic particles that have
sufcient surface area to sustain the cathodic reactions [41]. The
above postulations are going to be investigated by SEM examination
in Section 3.3.2.

1000
500
0
-500
-1000
-1500
-2000

0.00001

0.0001

0.001

0.01

0.1

Current density (mA/cm 2)


Fig. 12. Cyclic potentiodynamic polarization curves of monolithic and reinforced
materials at various combinations of carbide reinforcement in terms of carbide type
and volume fraction (DHS, 25 C). (a) Variation of carbide type (1.0 vol%), (b)
variation of TiC volume fraction, (c) variation of WC volume fraction, and (d)
variation of carbide type (0.7 vol% TiC, 0.5 vol% WC).

Al-Fe based intermetallic compounds are nobler than their Al


matrix [89]. At the pH of the electrolyte (pH  5.0), both AlFe
intermetallics and Al display regions of passivity [92]. However,
the lm over an AlFe based intermetallic phase is thin and electronically conductive; thus, when in galvanic couple with Al, it still
has a proven ability to efciently sustain cathodic reactions [93].
Based on the above postulations, the following mechanism is
considered to have taken place:
(1) Once aggressive anion adsorption on the aluminium oxide
surface lm at the Al/AlFe intermetallic interface occurred,
an active centre was developed. The active centre was then
the site for accelerated lm thinning [88].

1132

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

Potential (mV, Ag/AgCl)

-300

AlFe or Al and AlFeSi (even with the lm on the intermetallic phase), localized dissolution of the anodic Al (in or
adjacent to the eutectic microconstituent) occurred and
small pits were formed.
(3) As the pits were getting deeper, differential aeration cells
were formed between the bottom of the pits and the pit
walls.
(4) Pitting was evolved to intergranular corrosion at the boundaries where the Al-Fe intermetallics exist.

-500
-700
-900
-1100
-1300
-1500
-1700
0.0001

Forward Al 1050
Forward Al-1.0 vol%TiC
Forward Al-1.0 vol%WC

0.001

0.01

0.1

10

Current density (mA/cm 2)

Potential (mV, Ag/AgCl)

-300

-500
-700
-900
-1100

On closer inspection of Fig. 14, the following observations are


made: (a) features associated with the carbide reinforcement (clusters of aluminides and carbides) along with their Al-matrix have
remained intact of corrosion; (b) often, aluminide and carbide clusters have inhibited IC progress by acting as physical barriers; (c)
furthermore, in Fig. 14c, the voids around Al12W particles indicate
that the latter have acted as large cathodic sites.
The microstructural observations correlate well with the potentiodynamic polarization performance of the reinforced materials,
with respect to the following aspects:

-1300
-1500
-1700
0.0001

(a) The fact that corrosion was mainly associated with the intergranular iron aluminide phase (a feature of the monolithic
alloy), whilst features associated with the carbide reinforcement have remained essentially free of corrosion traces support the claim deriving from the similar polarization
behaviour in Fig. 12 and formulated in Section 3.3.1: the
corrosion behaviour of the tested materials was mainly
controlled by the corrosion of the monolithic alloy.
(b) The only notable evidence of corrosion associated with the
reinforcement is the anodic dissolution of Al around A12W
indicating a cathodic role for the large surface of Al12W
particles. This evidence correlates well with the relatively
high cathodic currents recorded during polarization of the
WC reinforced AMCs (in comparison with the TiC-AMCs,

Forward Al 1050
Forward Al-0.5 vol%WC
Forward Al-1.0 vol%WC

0.001

0.01

0.1

10

Current density (mA/cm 2)


Fig. 13. Cathodic polarization curves of monolithic and reinforced materials (DHS,
25 C). (a) Variation of carbide type (1.0 vol%), and (b) variation of WC volume
fraction.

(2) Once the lm was sufciently thinned, direct attack of the


exposed metallic Al occurred. Because the lm was thinned
locally, the attack on the metal was also concentrated. In this
stage, the electrochemical behaviour of the iron aluminide
became of major importance. Consequently, due to the
electrochemical potential difference between Al and

1.0 vol% TiC

1.0 vol% WC

1.0 vol% WC

1.0 vol% WC

Fig. 14. SEM micrographs of AMCs after cyclic polarization in DHS, 25 C (cross-sections). (a) and (b): extensive intergranular corrosion associated with the AlFe/Al and
AlFeSi/Al eutectic microconstituent; (c) evidence that Al12W particles have acted as cathodic surfaces. (d) WC & aluminide clusters and their matrix appear corrosion-free.
Black outlined ellipses: intergranular clusters of aluminides (Al5W, Al3Ti)/carbides and their matrix have remained free of corrosion signs. White arrows: aluminide (Al5W,
Al3Ti) and carbide clusters have acted as physical barriers to IC progress. White outlined ellipses: clusters of aluminides/carbides and their matrix have remained free of
corrosion signs.

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

Fig. 12a and d and Fig. 13a) and the 1.0 vol% WC-AMC (in
comparison with the 0.5 vol% WC-AMC, Figs. 12c and 13b).
According to Birbilis and Buchheit [94], the size of intermetallic particles plays an important role on the kinetics of
cathodic reactions, as this will govern the amount of current
the intermetallic can support.

1133

 Overall, the addition of submicron reinforcing carbide particles


had a benecial effect on the wear response of the monolithic
matrix, whilst it did not worsen the corrosion response irrespective of the particle volume fraction.

Acknowledgements
It should also be noted, that although clusters of aluminide particles and carbide nanocores do not appear to have acted as cathodic sites (even in the case of large surface area occupation, as in
Fig. 14ac), it is reasonable to assume that the entire volume of
high conductivity may have also contributed a cathodic inuence
on the matrix in compatibility with Deuis et al. [95].
The above observations suggest that the matrix/primary reinforcement interface was not the main factor in the corrosion
behaviour of these composites. The fact that the Al/WC and the
Al/TiC interfaces have remained unaffected in conjunction with
the sliding wear response, indicate the existence of clean and
strong reinforcement-matrix interfaces.

4. Conclusions
Aluminium Matrix Composites (AMCs) were produced by the
addition of sub-micron sized WC and TiC particles (61.0 vol%) into
a melt of Al1050. Casting was assisted by the employment of K2TiF6
as a wetting agent and mechanical stirring. The main conclusions
drawn from the study of the microstructure, wear performance
and corrosion performance of the cast materials are:
 The AMC contained two main types of reinforcement: (a) in-situ
Al3Ti and (Al5W + Al12W + Al3(Ti,W)) intermetallic particles in
the cases of TiCp-AMC and WCp-AMC, respectively; (b) carbide
nanoparticles, as such, or more often as clusters of carbide
nanocores and aluminide particles.
 Particle distribution was considered as reasonably uniform
comprising both clusters and isolated particles. Carbide nanocores, in clusters with aluminides or isolated were mainly
located at grain boundaries and the areas of the lastly solidied
liquid.
 The sliding wear performance of the alloy was markedly
improved by the addition of the carbide phase through a direct
effect stemming from the TiC and WC particles, as such and as
clusters with intermetallic phase, and an indirect effect originating from the intermetallic hard particles.
 WCp-AMCs presented higher wear resistance than TiCp-AMCs.
 A wear mechanism was formulated including four steps: (a)
plastic deformation of the Al-matrix leading to a repeated
hill-valley morphology of the wear surface; (b) formation
of surface alumina based oxide layers due to frictional heating; (c) crack formation due to friction, fatigue and brittleness
of the oxide surface layers; (d) ne groove formation along
the sliding direction. The wear mode was characterized as
mild.
 The corrosion behaviour of the reinforced materials in DHS,
at 25 C, was mainly controlled by the corrosion of the
alloy matrix. As such, the predominating form of corrosion
was intergranular corrosion (IC) of Al in the Al/iron aluminide eutectic microconstituent or adjacent to the grain
boundaries.
 Features associated with carbide reinforcement (clusters of
aluminide and carbide nanoparticles) along with their Almatrix have remained intact of corrosion, whilst in several
cases, they have acted as physical barriers to the IC progress.
However, a cathodic action of large Al12W particles was
observed.

To the Greek General Secretariat for Research & Technology and


the European Commission (NSRF 2007-2013) within the framework of Joint Research & Technology Programs /Hellas-Hungary.
References
[1] Kainer KU. Basics of metal matrix composites. In: Kainer KU, editor. Metal
matrix composites: custom-made materials for automotive and aerospace
engineering. Weinheim: Wiley-VCH Verlag GmbH & Co; 2006. p. 152.
[2] Surappa MK. Aluminium matrix composites: challenges and opportunities.
Sadhana 2003;28:31934.
[3] Surrapa MK, Rohatgi PK. Preparation and properties of cast aluminium ceramic
particle composites. J Mater Sci 1981;16:98393.
[4] Karbalaei Akbari M, Mirzaee O, Baharvandi HR. Fabrication and study on
mechanical properties and fracture behavior of nanometric Al2O3 particlereinforced A356 composites focusing on the parameters of Vortex method.
Mater Des 2013;46:199205.
[5] Ahamed H, Senthilkumar V. Experimental investigation on newly developed
ultrane-grained aluminium based nano-composites with improved
mechanical properties. Mater Des 2012;37:18292.
[6] Mula S, Padhi P, Panigrahi SC, Pabi SK, Ghosh S. On structure and mechanical
properties of ultrasonically cast Al-2% A1203 nanocomposite. Mater Res Bull
2009;44:115460.
[7] Haoze Liu, Luhong Wang, Aimin Wang, Taiping Lou, Bingzhe Ding, Zuangqi Hu.
Studies of SiC/Al nanocomposites under high pressure. Nanostruct Mater
1997;9:2258.
[8] Ezatpour HR, Sajjadi SA, Sabzevar MH, Huang Yizhong. Investigation of
microstructure and mechanical properties of Al6061-nanocomposite
fabricated by stir casting. Mater Des 2014;55:9218.
[9] Tahamtan S, Halvaee A, Emamy M, Zabihi MS. Fabrication of Al/A206-Al2O3
nano/micro composite by combining ball milling and stir casting technology.
Mater Des 2013;49:34759.
[10] Su H, Gao W, Feng Z, Lu Z. Processing, microstructure and tensile properties of
nano-sized Al2O3 particle reinforced aluminum matrix composites. Mater Des
2012;36:5906.
[11] Sharitabar M, Sarani A, Khorshahian S, Shaee Afarani M. Fabrication of
5052Al/Al2O3 nanoceramic particle reinforced composite via friction stir
processing route. Mater Des 2011;32:416472.
[12] Mazahery A, Abdizadeh H, Baharvandi HR. Development of high-performance
A356/nano-Al2O3 composites. Mater Sci Eng A 2009;518:614.
[13] Dehghan Hamedan A, Shahmiri M. Production of A356-1 wt% SiC
nanocomposite by the modied stir casting method. Mater Sci Eng A 2012;
556:9216.
[14] Alizadeh A, Taheri-Nassaj E, Hajizamani M. Hot extrusion process effect on
mechanical behavior of stir cast Al based composites reinforced with
mechanically milled B4C nanoparticles. J Mater Sci Technol 2011;27:11139.
[15] Wang J, Yi D, Su X, Yin F, Li H. Properties of submicron A1N particulate
reinforced aluminum matrix composite. Mater Des 2009;30:7881.
[16] Ansary Yar A, Montazerian M, Abdizadeh H. Microstructure and mechanical
properties of aluminum alloy matrix composite reinforced with nano-particle
MgO. J Alloys Compd 2009;484:4004.
[17] Naidich JV. The wettability of solids by liquid metals. Prog Surf Membr Sci
1998;14:353492.
[18] Mavros H, Karantzalis AE, Lekatou A. Solidication observations and sliding
wear behaviour of cast TiC particulate reinforced AlMgSi matrix composites. J
Compos Mater 2013;47:214962.
[19] Gopalakrishnan S, Murugan N. Production and wear characterization of AA
6061 matrix titanium carbide particulate reinforced composite by enhanced
stir casting method. Composites Part B 2012;43:3028.
[20] Kaftelen H, nl N, Gller G, Lt veoglu M, Henein H. Comparative
processing-structureproperty studies of AlCu matrix composites reinforced
with TiC particulates. Composites Part A 2011;42:81224.
[21] Karantzalis AE, Lekatou A, Georgatis E, Poulas V, Mavros H. Microstructural
observations of cast AlSiCuTiC composite material. J Mater Eng Perform
2010;19:58590.
[22] Peijie Li, Kandalova EG, Nikitin VI. In situ synthesis of AlTiC in aluminum
melt. Mater Lett 2005;59:25458.
[23] Contreras A, Bedolla E, Prez R. Interfacial phenomena in wettability of TiC by
AlMg alloys. Acta Mater 2004;52:98594.
[24] Lopez VH, Kennedy AR. Flux-assisted wetting and spreading of Al on TiC. J
Colloid Interface Sci 2006;298:35662.
[25] Karantzalis AE, Lekatou A, Georgatis E, Tsiligiannis Th, Mavros H. Solidication
observations in dendritic cast Al-alloys reinforced with TiC particles. J Mater
Eng Perform 2010;19:126875.

1134

A. Lekatou et al. / Materials and Design 65 (2015) 11211135

[26] Baumli P, Sychev J, Budai I, Szabo JT, Kaptay G. Fabrication of carbon ber
reinforced aluminum matrix composites via titanium-ion containing ux.
Composites Part A 2013;44:4750.
[27] Rajan HB, Ramabalan S, Dinaharan I, Vijay SJ. Synthesis and characterization of
in situ formed titanium diboride particulate reinforced AA7075 aluminum
alloy cast composites. Mater Des 2013;44:43845.
[28] Kalaiselvan K, Murugan N, Parameswaran S. Production and characterization
of AA6061-B4C stir cast composite. Mater Des 2011;32:40049.
[29] Toptan F, Kilicarslan A, Karaaslan A, Cigdem M, Kerti I. Processing and
microstructural characterization of AA 1070 and AA 6063 matrix B4Cp
reinforced composites. Mater Des 2010;31:S8791.
[30] De Salazar JMG, Urena A, Manzanedo S, Barrena MI. Corrosion behavior of
AA6061 and AA7005 reinforced with Al2O3 particles in aerated 3.5% chloride
solutions: potentiodynamic measurements and microstructure evaluation.
Corros Sci 1999;41:52945.
[31] Shimizu Y, Nishimura T, Matsushima I. Corrosion resistance of Al-based metal
matrix composites. Mater Sci Eng A 1995;198:1138.
[32] Buarzaiga MM, Thorpe SJ. Corrosion behaviour of as-cast silicon carbide
particulate aluminium alloy metal-matrix composites. Corrosion Metals Alloys
1994;50:17685.
[33] Trzaskoma PP. Pit morphology of aluminum alloy and silicon carbide/
aluminum alloy metal matrix composites. Corrosion Metals Alloys 1990;46:
4029.
[34] Mclntyre JF, Conrad RK, Golledge SL. Technical note: the effect of heat
treatment on the pitting behavior of SiCW/AA2124. Corrosion 1990;46:9025.
[35] Zhu J, Hihara LH. Corrosion of continuous alumina-bre reinforced Al-2 wt.%
CuT6 metalmatrix composite in 3.15 wt.% NaCl solution. Corros Sci 2010;52:
40615.
[36] Nath D, Namboodhirt TKG. Some corrosion characteristics of aluminium-mica
particulate composites. Corros Sci 1989;29:121521.
[37] Hihara LH, Latanision RM. Galvanic corrosion of aluminium-matrix
composites. Corrosion 1992;48:54652.
[38] Trowsdale AJ, Noble B, Harris SJ, Gibbins ISR, Thompson GE, Woods GC. The
inuence of silicon carbide reinforcement on the pitting behaviour of
aluminium. Corros Sci 1996;38:17791.
[39] Paciej RC, Agarwala VS. Inuence of processing variables on the corrosion
susceptibility of metal-matrix composites. Corrosion 1998;44:6804.
[40] Pardo A, Merino MC, Merino S, Viejo F, Carboneras M, Arrabal R. Inuence of
reinforcement proportion and matrix composition on pitting corrosion
behaviour of cast aluminium matrix composites (A3xx.x/SiCp). Corros Sci
2005;47:175064.
[41] Grifths AJ, Turnbull A. An investigation of the electrochemical polarization
behaviour of 6061 aluminum metal matrix composites. Corros Sci
1994;36:2335.
[42] Kiourtsidis GE, Skolianos SM, Pavlidou EG. A study on pitting behaviour of
AA1913: SiCp composites using the double cycle polarization technique.
Corros Sci 1999;41:1183203.
[43] Alaneme KK, Bodunrin MO. Corrosion behavior of alumina reinforced
aluminum (6063) metal matrix composites. J Miner Mater Character Eng
2011;10:115365.
[44] Candan S. An investigation on corrosion behaviour of pressure inltrated Al
Mg alloy/SiCp composites. Corros Sci 2009;51:13928.
[45] Seah KHW, Krishna M, Vijayalakshmi VT, Uchil J. Corrosion behaviour of garnet
particulate reinforced LM13 Al alloy MMCs. Corros Sci 2002;44:91725.
[46] Jameel AA, Nagaswarupa HP, Krupakara PV, Vijayamma KC. Corrosion
characterization of Al 6061/Zircon metal matrix composites in acid chloride
mediums by open circuit potential studies. Int J Appl Chem 2009;5:110.
[47] Nagaswarupa HP, Bheemanna HG, Banuprakash G. Electrochemical studies of
Al 7075/zircon metal matrix composites in natural sea water. Ultra Chem
2012;8:31928.
[48] Toptan F, Alves AC, Kerti I, Ariza E, Rocha LA. Corrosion and tribocorrosion
behaviour of AlSiCuMg alloy and its composites reinforced with B4C
particles in 0.05 M NaCl solution. Wear 2013;306:2735.
[49] Bindumadhavan PN, Wah HK, Prabhaka O. Dual particle size (DPS) composites:
effect on wear and mechanical properties of particulate metal matrix
composites. Wear 2001;248:11220.
[50] Mandal A, Murty BS, Chakraborty M. Sliding wear behaviour of T6 treated
A356-TiB2 in-situ composites. Wear 2009;266:86572.
[51] Yang LJ. The effect of nominal specimen contact area on the wear coefcient of
A6061 aluminium matrix composite reinforced with alumina particles. Wear
2007;263:93948.
[52] Kk M, zdin K. Wear resistance of aluminium alloy and its composites
reinforced by Al2O3 particles. J Mater Process Technol 2007;83:3019.
[53] Yalcin Y, Akbulut H. Dry wear properties of A356-SiC particle reinforced MMCs
produced by two melting routes. Mater Des 2006;27:87281.
[54] Deuis RL, Subramanian C, Yellup JM. Dry sliding wear of aluminium
composites a review. Compos Sci Technol 1997;57:41535.
[55] Sannino AP, Rack HJ. Dry sliding wear of discontinuously reinforced
aluminium composites: review and discussion. Wear 1995;189:119.
[56] Veeresh Kumar GB, Rao CSP, Selvaraj N. Mechanical and tribological behavior
of particulate reinforced aluminum metal matrix composites a review. J
Miner Mater Character Eng 2011;10:5991.
[57] Suh NP. The delamination theory of wear. Wear 1973;25:11124.
[58] Archard JF. Contact and rubbing of at surfaces. J Appl Phys 1953;24:9818.
[59] Heilmann P, Don J, Sun TC, Rigney DA. Sliding wear and transfer. Wear
1983;91: 17190.

[60] Al-Qutub AM, Allam IM, Qureshi TW. Effect of sub-micron Al2O3 concentration
on dry wear properties of 6061 aluminum based composite. J Mater Process
Technol 2006;172:32731.
[61] Karbalaei Akbari M, Baharvandi HR, Mirzaee O. Nano-sized aluminum oxide
reinforced commercial casting A356 alloy matrix: evaluation of hardness,
wear resistance and compressive strength focusing on particle distribution in
aluminum matrix. Composites Part B 2013;52:2628.
[62] Mohammad Shari E, Karimzadeh F, Enayati MH. Fabrication and evaluation of
mechanical and tribological properties of boron carbide reinforced aluminum
matrix nanocomposites. Mater Des 2011;32:326371.
[63] Sameezadeh M, Emamy M, Farhangi H. Effects of particulate reinforcement
and heat treatment on the hardness and wear properties of AA 2024-MoSi2
nanocomposites. Mater Des 2011;32:215764.
[64] Anya CC. Wet erosive wear of alumina and its composites with SiC nanoparticles. Ceram Int 1998;24:53342.
[65] Ravindran P, Manisekar K, Vinoth Kumar S, Rathikac P. Investigation of
microstructure and mechanical properties of aluminum hybrid nano
composites with the additions of solid lubricant. Mater Des 2013;51:
44856.
[66] Nemati N, Khosroshahi R, Emamy M, Zolriasatein A. Investigation of
microstructure, hardness and wear properties of Al-4.5 wt.% CuTiC
nanocomposites produced by mechanical milling. Mater Des 2011;32:
371829.
[67] Battocchi D, Simes AM, Tallman DE, Bierwagen GP. Comparison of testing
solutions on the protection of Al-alloys using a Mg-rich primer. Corros Sci
2006;48:222640.
[68] Witusiewicz VT, Bonder AA, Hecht U, Rex S, Velikanova TYa. The AlBNbTi
system III. Thermodynamic re-evaluation of the constituent binary system Al
Ti. J Alloys Compd 2008;465:6477.
[69] Franke P, Neuschtz D, editors. Elements and binary systems from AgAI to
AuTI, series: Landolt-Brnstein: numerical data and functional relationships
in science and technology New series, Part 19B1, Subseries: Physical
Chemistry. Berlin, Heidelberg: Springer; 2002. p. 2146.
[70] NPL (National Physical Laboratory), MTDATA Phase Diagram Software, Ti-W
phase diagram, calculated 15-07-2003, <http://resource.npl.co.uk/mtdata/
phdiagrams/tiw.htm>.
[71] NPL (National Physical Laboratory), MTDATA Phase Diagram Software, AlC
phase diagram, calculated 09-03-1999, <http://resource.npl.co.uk/mtdata/
phdiagrams/png/alc.png>.
[72] El-Mahallawy N, Taha MA, Jarfors AEW, Fredriksson H. On the reaction
between aluminium, K2TiF6 and KBF4. J Alloys Compd 1999;292:2219.
[73] Birol Y. Production of AlTiB master alloys from Ti sponge and KBF4. J Alloys
Compd 2007;440:10812.
[74] Contreras A, Angeles-Chvez C, Flores O, Perez R. Structural, morphological
and interfacial characterization of AlMg/TiC composites. Mater Character
2007;58:68593.
[75] House JE. Inorganic chemistry. Academic Press; 2012. p. 353.
[76] Jahntek M, Kraj M, Hafner J. Interatomic bonding, elastic properties, and
ideal strength of transition metal aluminides: a case study for Al3(V,Ti). Phys
Rev B 2005;71. 024101(1-16).
[77] Lide DR, editor. CRC Handbook of Chemistry and Physics. 72nd ed. CRC Press;
19911992.
[78] Georgatis E, Lekatou A, Karantzalis AE, Petropoulos H, Katsamakis S, Poulia
A. Development of a cast AlMg2SiSi in-situ composite: microstructure,
heat treatment and mechanical properties. J Mater Eng Perform 2013;22:
72941.
[79] Zubko AM, Lobanov VG, Nikonova VV. Reaction of foreign particles with
crystallization front. Sov Phys Crystallogr 1978;18:23941.
[80] Tyagi R. Synthesis and tribological characterization of in situ cast AlTiC
composites. Wear 2005;259:56976.
[81] Ranganath G, Sharma SC, Krishna M. Dry sliding wear of garnet reinforced
zinc/aluminium metal matrix composites. Wear 2001;251:140813.
[82] Venkataraman B, Sundararajan G. The sliding wear behaviour of AlSiC
particulate composites I. Macrobehaviour. Acta Mater 1996;44:
45160.
[83] Sannino AP, Rack HJ. Tribological investigation of 2009 Al-20 vol% SiCp/17-4
PH Part I: composite performance. Wear 1996;197:1519.
[84] Sarkar AD. Friction and wear. London: Academic Press; 1980. p. 2059.
[85] Urena A, Rams J, Campo M, Snchez M. Effect of reinforcement coatings on the
dry sliding wear behaviour of aluminium/SiC particles/carbon bres hybrid
composites. Wear 2009;266:112836.
[86] Las L, Rodrigez-Ibabe JM. Wear behaviour of eutectic and hypereutectic AlSi
CuMg casting alloys tested against a composite brake pad. Mater Sci Eng A
2003;363:193200.
[87] Niu Haiyang, Chen Xing-Qiu, Liu Peitao, Xing Weiwei, Cheng Xiyue, Li
Dianzhong, et al. Extra-electron induced covalent strengthening and
generalization of intrinsic ductile-to-brittle criterion. Sci Rep 2012;2:718.
10.1038/srep0071.
[88] Foley RT. Localized corrosion of aluminum alloys a review. Corrosion 1986;
42:27788.
[89] Szklarska-Smialowska Z. Pitting corrosion of aluminum. Corros Sci 1999;41:
174367.
[90] Holm K, Hornbogen E. Annealing of supersaturated and deformed Al0.042 wt% Fe solid solutions. J Mater Sci 1970;5:65562.
[91] Stickels CA, Bush RH. Precipitation in the System Al0.05 wt pct Fe. Metall Trans
1971;2:203142.

A. Lekatou et al. / Materials and Design 65 (2015) 11211135


[92] Birbilis N, Buchheit RG. Investigation and discussion of characteristics for
intermetallic phases common to aluminum alloys as a function of solution pH.
J Electrochem Soc 2008;155(3):C11726.
[93] Stansbury EE, Buchanan RA. Fundamentals of electrochemical corrosion.
Materials Park Ohio: ASM Int.; 2000. p. 326.

1135

[94] Birbilis N, Buchheit RG. Electrochemical characteristics of intermetallic phases


in aluminum alloys: an experimental survey and discussion. J Electrochem Soc
2005;152(4):B14051.
[95] Deuis RL, Green L, Subramanian C, Yellup JM. Corrosion behaviour of
aluminium composite coatings. Corrosion 1997;53:88090.

Você também pode gostar