Materials Science & Engineering A: Fei Chen, Zongning Chen, Feng Mao, Tongmin Wang, Zhiqiang Cao

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Materials Science & Engineering A 625 (2015) 357368

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

TiB2 reinforced aluminum based in situ composites fabricated


by stir casting
Fei Chen a, Zongning Chen b, Feng Mao a, Tongmin Wang a,n, Zhiqiang Cao b,nn
a
Key Laboratory of Materials Modication by Laser, Ion, and Electron Beams (Ministry of Education), School of Materials Science and Engineering,
Dalian University of Technology, Dalian 116024, China
b
Laboratory of Special Processing of Raw Materials, Dalian University of Technology, Dalian 116024, PR China

art ic l e i nf o

a b s t r a c t

Article history:
Received 22 August 2014
Received in revised form
4 December 2014
Accepted 8 December 2014
Available online 18 December 2014

In this study, a new technique involving mechanical stirring at the salts/aluminum interface was
developed to fabricate TiB2 particulate reinforced aluminum based in situ composites with improved
particle distribution. Processing parameters in terms of stirring intensity, stirring duration and stirring
start time were optimized according to the microstructure and mechanical properties evaluation. The
results show that, the rst and last 15 min of the entire 60 min holding are of prime importance to the
particle distribution of the nal composites. When applying 180 rpm (revolutions per minute) stirring at
the salts/aluminum interface in these two intervals, a more uniform microstructure can be achieved and
the Al-4 wt% TiB2 composite thus produced exhibits superior mechanical performance. Synchrotron
radiation X-ray computed tomography (SR-CT) was used to give a full-scale imaging of the particle
distribution. From the SR-CT results, the in situ AlxTiB2 composites (x 1, 4 and 7, all in wt%) fabricated
by the present technique are characterized by ne and clean TiB2 particles distributed uniformly
throughout the Al matrix. These composites not only have higher yield strength (0.2) and ultimate
tensile strength (UTS), but also exhibit superior ductility, with respect to the AlTiB2 composites
fabricated by the conventional process. The 0.2 and UTS of the Al7TiB2 composite in the present work,
are 260% and 180% higher than those of the matrix. A combined mechanism was also presented to
interpret the improvements in yield strength of the composites as inuenced by their microstructures
and processing history. The predicted values are in good agreement with the experimental results,
strongly supporting the strengthening mechanism we proposed. Fractography reveals that the composites thus fabricated, follow ductile fracture mechanism in spite of the presence of stiff reinforcements.
& 2014 Elsevier B.V. All rights reserved.

Keywords:
TiB2 particles
In situ composites
Stirring
Tomography
Microstructure
Mechanical properties

1. Introduction
Al based metal matrix composites (MMCs) reinforced with
ceramic particles have received extensive attention due to their
high specic strength-to-weight ratio, good wear resistance,
excellent dimensional stability and superior damping capacity in
comparison with the matrix alloy [1,2]. The conventional practice
to prepare Al based composites (ex situ composites) involves the
addition of externally synthesized reinforcements, such as SiC,
Al2O3 and TiC, to the matrix alloys [35]. This process, however,
could lead to segregation and thermodynamic instability of the
reinforcements, and poor adhesion at the interface, unless the
ceramic particles have been suitably modied [6,7].

Corresponding author. Tel.: 86 411 84706790; fax: 86 41184706790.


Corresponding author. Tel.: 86 411 84706169; fax: 86 411 8470616.
E-mail addresses: [email protected] (T. Wang), [email protected] (Z. Cao).

nn

http://dx.doi.org/10.1016/j.msea.2014.12.033
0921-5093/& 2014 Elsevier B.V. All rights reserved.

To overcome these drawbacks often occasioned in the preparation of ex situ MMCs, in situ techniques have been greatly
developed in recent years. Since the formation of the reinforcements takes place within the matrix, in situ synthesized MMCs
provide advantages including uniform distribution of ner particles, excellent bonding at the matrix/reinforcement interface,
thermodynamical stability of the reinforcements and process
economy [8,9].
A wide variety of in situ formed ceramic particulates, such as
Al2O3, TiB2 and TiC, have been used as reinforcements to fabricate
Al based MMCs [10,11]. Among these particulates, TiB2 is an
advanced strengthening phase for Al matrix as it possesses a
desirable combination of physical and mechanical properties,
including high melting point (3225 1C), high elastic modulus
(534 GPa), high hardness (3400 HV) and outstanding wear resistance. More importantly, it does not react with Al to form any
detrimental reaction products at the matrix/reinforcement interface [7,12]. In the past two decades, numerous studies have been
conducted to develop new fabrication processes of TiB2 reinforced

358

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

in situ Al based MMCs with improved wear performance, mechanical properties and damping capacity [6,13,14].
Although the TiB2 reinforced in situ Al based composites have
prominent advantages, their microstructure and mechanical properties are known to be highly sensitive to the processing parameters in the production. Reinforcement segregation and residual
intermediate products, such as TiAl3 and AlB2, are frequently
present in the nal composites [1517]. These defects have been
found to act as preferential sites for crack initiation and propagation. Therefore, elimination of these defects is regarded as a
primary concern to prepare AlTiB2 composites with consistent
properties [18,19].
In the present study, the stir casting technique, a conventional
method to prepare ex situ MMCs [20,21], was elaborately modied
to fabricate TiB2 reinforced in situ Al based MMCs using the halide
salt route. In this process, a well-designed device was applied to
stirring at the salts/aluminum interface, based on the fact that the
reactions, by which TiB2 is formed, generally take place in the
vicinity of the salts/aluminum interface. The effects of processing
parameters, i.e. stirring speed, stirring duration, and stirring start
time on the microstructure and mechanical properties of the nal
products have been studied in detail.
Furthermore, synchrotron radiation X-ray computed tomography (SR-CT) has been used to investigate the particle distribution
of the experimental AlTiB2 composites from three dimensional
(3D) view, since the third-generation synchrotron radiation (SR)
source has been reported being able to offer a unique opportunity
to non-destructively observe the microstructural characteristics of
metallic materials with submicron spatial resolution [22,23].
However, in the conventional 2D (two dimensional) measurements, such as, optical and scanning electron microscope analysis,
materials microstructure is likely to be destroyed during sample
preparation (cutting, polishing and etching) [2426]. The aim of
the present work is to explore the feasibility of fabricating low cost
Al based composites with improved mechanical properties, in
addition, to enlarge the application of synchrotron radiation in
the eld of MMCs.

2. Experimental procedures
AlTiB2 in situ composites were synthesized by the exothermic
reaction of halide salts (KBF4K2TiF6) with molten aluminum
according to the following reactions:
3K2TiF6 13Al-3KAlF4 K3AlF6 3TiAl3

(1)

2KBF4 3Al-2KAlF4 AlB2

(2)

TiAl3 AlB2-TiB2 4Al

(3)

Commercially available Al (99.8% Al: all compositions are in


mass fraction unless stated otherwise), KBF4 and K2TiF6 powders
of analytical purity (99% purity) were used as starting materials.
Fig. 1 shows the schematic diagram of the set-up for fabricating
AlTiB2 in situ composites. For each experiment, 1.5 kg Al was
melted and heated to 860 75 1C in a graphiteclay crucible under
ambient atmospheres, using a resistance furnace. The pre-dried
(200 1C for 2 h) K2TiF6 and KBF4 mixture with a stoichiometric Ti/B
ratio of 1:2 was then added into the melt. After salts melted, a
preheated impeller with four straight blades was employed to
stirring at the salts/aluminum interface. The straight blades of the
impeller, which could reduce the inclusion of post-reaction slag,
were made of graphite due to the fact that graphite will not
contaminate both of the molten salts and molten aluminum.
Moreover, it has been demonstrated that carbon plays a positive
role in the formation of TiB2.

Fig. 1. Schematic diagram of the experimental set-up for fabricating AlTiB2 in situ
composites.

Fig. 2. Schematic diagram of experimental setup of synchrotron radiation X-ray


computed tomography.

The stirring intensity (60, 180 and 300 rpm (revolutions per
minute)), stirring duration (5, 15, 30 and 60 min) and stirring start
time (initial, middle and later) were varied to investigate the
effects of stirring parameters on the microstructures and mechanical performances of the Al4% TiB2 composites. The melt was held
at 860 1C for 60 min to enable the reactions to achieve completion.
After decanting by-product KAlF salts and degassing for one
minute by high-purity argon, the melt of 700 1C was poured into a
preheated (200 1C) permanent mold. Composites with 1% and 7%
TiB2 were also fabricated under the optimized process condition to
investigate the feasibility of the technique in producing composites with different particle contents.
Phase identication of the composites was performed by an
X-ray diffractometer (Empyrean) using Cu-K radiation operated
at 40 kV and 40 mA. For microstructure analysis, the specimens
were polished and etched with Keller's reagent, then examined by
a eld emission scanning electron microscope (SEM) equipped
with an energy dispersive spectrometer (EDS).
The synchrotron radiation analysis was carried out at beam line
BL13W1 at Shanghai Synchrotron Radiation Facility (SSRF). Experimental setup is shown in Fig. 2. Samples with 0.7 mm in diameter
and 20 mm in length were cut from ingots contain 7% TiB2 by
electro discharge machining. The sample was xed upright with
its axis perpendicular to the incident monochromatic X-ray beam
and a beam of 18 keV was focused on and sent through the central
area of the sample. The transmitted beam was recorded by a CCD
based camera with a spatial resolution of 0.37 m placed 45 cm
behind the specimens. Phase contrast (sensitive to internal structure change), and absorbing contrast (sensitive to density difference) were used to establish images. During the experiment, the
sample rotated about the vertical axis and the camera collected

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

the image data continuously with exposure time per frame was
2 s. A total of 1048 images were used to reconstruction three
dimensions (3D) structure of the composites.
Hardness tests of the composites were carried out using a
Brinell hardness tester. The tensile specimens with a gauge length
of 30 mm, a gauge diameter of 6 mm were prepared according to
ASTM E8M-04 standard. The yield strength (0.2) and ultimate
tensile strength (UTS) were estimated using a computerized
universal testing machine. The tensile velocity was 0.05 mm/s.
Three tests were conducted for each composite to get a precise
value for each property. Fracture surfaces were analyzed using
SEM to evaluate the fracture mechanism.

3. Results and discussion


3.1. Stirring intensity
Five experiments summarized in Table 1 were carried out to
assess the effect of interface stirring intensity (0, 60, 180 and
300 rpm) on the quality of the Al  4% TiB2 in situ composites. The
XRD patterns of the composites are shown in Fig. 3. The main
phases in all the four composites are Al and TiB2. This is consistent
with the low Gibb's energy of TiB2 in comparison to AlB2 and TiAl3
[27], conrming the foregoing reactions (13) have achieved
completion. Contrary to the nding of Chen et al. [28] in the
present study, no trace of intermediate product TiAl3 or AlB2 can
be detected. Considering the TiB ratio is just 1:2, it is inferred that
no signicant KBF4 or K2TiF6 decomposed or oxide during the
fabrication process, regardless of whether applying salts/aluminum interface stirring or not. It should be note that, in the XRD
pattern of sample 4, apart from Al and TiB2, diffraction peaks of
K3AlF6 and Al2O3 also present, indicating that erce stirring leads

Table 1
Effect of stirring intensity on the mechanical properties of Al4% TiB2 in situ
composites.
Sample
no.

Stirring
intensity
(rpm)

Stirring
time
(min)

Hardness
(HB)

0.2 (MPa) UTS (MPa) Elongation


(%)

0a
1
2
3
4

0
0
60
180
300

0
0
15
15
15

18.2 7 0.7
23.4 7 1.8
23.6 7 1.1
24.47 0.3
24.27 0.6

24.67 2.7 48.47 3.3


46.17 3.8 79.17 4.6
55.6 7 3.5 93.2 7 3.7
62.3 7 1.9 104.47 2.4
60.2 7 2.2 101.5 7 2.6

39.3 7 3.2
25.2 7 4.1
34.4 7 2.8
35.67 2.1
32.3 7 3.2

0 stands for the pure Al used in the present work.

Fig. 3. XRD pattern of Al4% TiB2 in situ composites prepared with different
stirring intensities.

359

to some degree of inclusion. This is believed to be associated with


the deep vortex in the molten KAlF salts and the exposure of the
molten aluminum to atmospheric moisture.
Fig. 4 shows the microstructure and EDS spectrums of the
in situ composites fabricated with different stirring speed. It can
be seen that for the composite without stirring (sample 1),
particles distribution is quite non-uniform. Agglomerates in various
sizes (0.013 mm) and shapes for instants, string-like and irregular
shape, distribute in the matrix, accompanied by some coarse TiB2
block with size range from 10 to 40 m (Fig. 4(a) and (b)). These
defects may act as crack source during the loading of the composite.
After stirring for 15 min at a speed of 60 rpm (sample 2), the particle
distribution is improved (Fig. 4(c) and (d)), string-like agglomerates
disappear despite small agglomerates (1070 m) and coarse
TiB2 block still remained. TiB2 particles distribute non-uniformly
along the crystal boundary. As stirring speed increased to 180 rpm
(sample 3), TiB2 particles distribute more uniformly (Fig. 4(e) and (f)),
agglomerates and coarse TiB2 blocks observed in sample 1 and 2 are
no longer present, indicating the stirring intensity is strong enough
to suppress the formation of severe agglomeration.
Sample 4 exhibits preferred particulate features (Fig. 4(e) and
(f)), i.e. more homogeneous distribution and well developed
hexagonal morphology of TiB2 particles. However, the dirty
particle surface and undesirable oxides, veried by the blurred
particles and the above stated XRD patterns, frequently appear in
the composite.
In contrast to the coarse grain of unreinforced aluminum
(1077.1 m), all the composites have ner matrix grain, the average
grain size of samples 24 are 164.3, 140.9, 135.3 and 124.7 m,
respectively. This can be attributed to the grain renement of
the TiB2 particles. It is worth noting that, the grain size shows a
decline trend with the increasing in stirring speed, indicating the
good distribution of the particles is benecial to the grain
renement.
The mechanical performance shown in Table 1 reveal that all of
the four composites have superior hardness and strength with
respect to pure Al. The 0.2 and UTS increase with the increasing in
stirring speed up to 180 rpm. This is due to the two facts:
enhanced stirring intensity results in a more uniform particle
distribution; the average grain size of the composites decrease
with the increasing stirring speed. For stirring speed higher than
180 rpm, the degradation in the tensile properties is attributed to
the entrapped by-products and oxide inclusions.
3.2. Stirring duration
To investigate the effect of stirring duration on the nal quality
of AlTiB2 in situ composites, experiments with stirring duration
varied from 5 to 60 min and xed stirring speed of 180 rpm were
carried out (Table 2). The microstructure of the fabricated composites are shown in Fig. 5. The TiB2 distribution of the composite
with stirring for 5 min is inferior to that of sample 3, small
agglomerates (1050 m), coarse TiB2 blocks and rod-like titanium
boride gather in the grain boundaries (Fig. 5(a) and (b)). As the
stirring duration increased to 30 and 60 min (Fig. 5(e) and (h)),
TiB2 particles dominated in hexagonal and spherical morphology,
distribute homogeneously in the matrix. It is worth to note that,
no obvious change in particle morphology can be found among the
composites with stirring duration exceed 15 min, suggesting that
the particle morphology might be mainly determined in the initial
stage of fabrication process, while particle distribution affected by
the whole holding process. Thus, studying the effect of stirring at
different stages of fabricating process on the nal composite
microstructure and properties is necessary.
Table 2 reveals the tensile properties increase with the increasing in stirring duration. The UTS of sample 6 is 106.7 MPa, 1.2 times

360

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

Fig. 4. SEM micrographs of the Al4% TiB2 in situ composites prepared with different interface stirring intensity (a) and (b) non-stirring; (c) and (d) 60 rpm; (e) and (f)
180 rpm; (g) and (h) 300 rpm.

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

361

Table 2
Effect of stirring duration on the mechanical properties of Al4% TiB2 in situ composites.
Sample no.

Stirring intensity (rpm)

Stirring duration (min)

Hardness (HB)

0.2 (MPa)

UTS (MPa)

Elongation (%)

5
3
6
7

180
180
180
180

5
15
30
60

24.0 7 0.6
23.9 7 0.3
24.2 7 0.4
25.3 7 0.2

54.2 7 2.2
62.3 7 1.9
63.6 7 1.8
68.97 1.4

96.0 7 3.2
104.47 2.4
106.77 2.3
111.4 7 1.7

34.2 7 2.8
35.6 7 2.1
35.8 7 1.7
36.3 7 1.3

higher than that of the matrix. As stirring duration increased to


60 min, UTS is further improved. Since no obviously difference in
particle morphology can be observed, the further enhancement in
the properties is mainly attributed to the improved particle
distribution. In addition, the decreased grain size, which reduces
from 143.7 for sample 5 to 128.4 and 103.2 m for sample 6 and
7 respectively, also accounts partially for the improvement in the
tensile strength.
3.3. Stirring at different stages of the fabrication process
To gure out the effect of the interface stirring at different
stages of the fabrication process on the nal composite microstructure and properties, another two experiments (Table 3)
involving starting the stirring at different time after salts melted
(t0) were carried out to compare with sample 3 (initial stirring).
SEM micrographs of the three composites samples are shown
in Fig. 6. The composite with middle stirring (sample 8) exhibits
that a few agglomerates are remained and main particles segregate in the grain boundaries. However, a few particles disperse in
the inner-grain (Fig. 6(c) and (d)). With effect similar to, but more
pronounced than that of the middle stirring, the later stirring
(sample 9) results in a more dispersed particle distribution, in
spite of many agglomerates still remained in the composite (Fig. 6
(e) and (f)). This suggests once these agglomerates are formed in
the molten matrix, it is difcult to be eliminated by the subsequent stirring owing to the high strength and good refractory of
TiB2.
Fig. 6(d) and (f) show the present of TiB2 particles with various
morphologies, including coarse block, rod like and rugby-shape,
implying both middle stirring and later stirring have little effect on
the growth of the particles, even though it could improve particle
dispersion. The comparison between the three samples conrms
the truth of the aforementioned deduction that the particle
morphology is mainly determined in the initial 15 min, while
particle distribution is affected by the whole holding process,
especially the last 15 min of the synthetic process.
The mechanical properties of the composites fabricated with
stirring at different process stage are shown in Table 3. As contrast
to the composite without stirring (sample 1), all the three
composites possess superior strength, indicating interface stirring
is benecial to the strength improvement regardless when it is
applied. Compared to the initial stirring and later stirring, middle
stirring results in a relative poor strength (Sample 8), which is
consistent with the relative coarse matrix grain (137.3 m), uneven
particle distribution, blocky and rod like particle morphology. It
should be note that, although the inferior particles morphology
and some agglomerates in the nal composite, sample 9 has a
tensile strength similar to that of sample 3. This can be ascribed to
the fact that more TiB2 particles well dispersing in the matrix as
well as the small grain size of 126.5 m.
3.4. Optimized fabricate process
Considering energy saving and the fact that the particle
morphology is mainly determined in the initial 15 min, while

particle distribution is affected by the whole holding process,


especially the last part of the synthetic process. A new stir-casting
technology, which involving the combination of the effect of the
initial stirring and the later stirring, i.e, 15 min initial stirring, after
that hold for 30 min, then stirring once again for 15 min up to
pouring, was adopted to obtain a more uniform particle distribution. Fig. 7 shows SEM images of the composite fabricated by the
developed twice interface stirring. It is obvious that, the composite
reveals a more uniform particle distribution.
Table 4 shows the mechanical properties of in situ composite
synthesized by the optimized fabrication process. The yield and
tensile strength of the composite increase by 66.0% and 134.7%
respectively as compared to those of the matrix, and are superior
to those of composites synthesized without stirring (sample 1) or
single stirring (sample 3, 8 and 9), even more, slightly higher than
that of the composite synthesized by long stirring of 60 min
(sample 7). This implies long continuous stirring may cause
entrapping of some inclusion. Based upon the above microstructure and mechanical properties analysis, it is concluded that
applying salts/aluminum interface stirring with speed of 180 rpm
in both the initial and the last 15 min of the 60 min composite
synthesis process, is a feasible process to synthesize dispersive
TiB2 particles reinforced Al4% TiB2 MMCs.

3.5. Investigating the effect of the salts/aluminum interface stirring


on the composites microstructure using SR-CT
Al7% TiB2 composite was fabricated to investigate the feasibility of the developed technique in high particle content composites producing. To comparison, the reference composite was
synthesized under the same process condition, except for no
stirring was applied. The 3D structure of the composites synthesized without and with the developed interface stirring were
reconstructed by SR-CT technique. Fig. 8(a) and (b) show the
views of the composite without stirring from different angles. It is
evident that the spatial distribution of the TiB2 particles is
remarkably non-uniform. Fine TiB2 particles segregate at grain
boundaries, together with some occulent and ake-shape
agglomerates and exceptional coarse particles. As can be seen
that, the occulent agglomerates which might result from the
emulsication of salt into the metal [29], have uncompacted
structure with size range from 50 to 300 m. Feng et al. reported
that string-like particle agglomerates are a distinct feature of the
Al  TiB2 in situ composites [30,31]. In the present work, although
string-like agglomerates can be found in SEM analysis, this is not
the case from 3D views. Instead, ake-shape agglomerates with
thickness of 515 m are the substitutes. This means that the real
morphology of common string-like agglomerates is ake-shape.
Fig. 8(c) shows the 3D image of the Al7% TiB2 composite
produced by the developed technique involving twice stirring at
salts/aluminum interface. It can be seen, upon mechanical stirring,
the situation is changed. The occulent and aky-shape particle
agglomerates, as well as the coarse TiB2 particles disappear, ne
TiB2 particles distribute homogeneously in the aluminum matrix.
This means that the developed salts/aluminum interface stirring

362

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

Fig. 5. SEM micrographs of the Al4% TiB2 in situ composites prepared with different stirring duration: (a) and (b) 5 min; (c) and (d) 15 min; (e) and (f) 30 min; (g) and (h)
60 min.

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

363

Table 3
Effect of stirring at different process stages on the mechanical properties of Al4% TiB2 in situ composites.
Sample no.

Stirring intensity (rpm)

Stirring start time (min)

Stirring duration (min)

Hardness (HB)

0.2 (MPa)

UTS (MPa)

Elongation (%)

3
8
9

180
180
180

t0
t0 15
t0 45

15
15
15

23.9 7 0.4
22.5 7 0.6
23.3 7 0.3

62.3 7 1.9
59.8 7 2.5
66.57 1.4

104.47 2.4
96.3 7 2.8
102.6 7 1.8

35.6 7 2.1
32.9 7 2.9
36.2 7 1.7

Fig. 6. SEM micrographs of the Al4% TiB2 composites fabricated by applying interface stirring at different time after salts melted: (a) t0 (initial stirring); (b) t0 15 (middle
stirring); (c) t0 45 min (later stirring).

technique is effective in synthesizing high content TiB2 reinforced


MMCs.
Interface mechanical stirring improves TiB2 particles distribution by two means. One is preventing the formation of various
agglomerates: enhancing diffusion at the salts/aluminum interface, not only suppresses the nucleation and growth of coarse AlB2
and TiAl3 intermetallic compounds, but also constrains the

formation and deposition of large engulf-salt droplets; in addition,


velocity gradient exists in the owing uid owning to the viscosity
of the melt, therefore, owing liquids scour the surface of existing
compounds and engulf-salt, accelerates the elements diffusion and
disperse TiB2 seeds into the bulk molten pool. The other effect is
resisting the sedimentation of the already formed particles. Twice
interface stirring, especially the later stirring, overcomes the

364

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

Fig. 7. SEM micrographs of the in situ Al4% TiB2 composite fabricated by the optimized fabricate process:( a) low magnicent; and (b) high magnicent.
Table 4
Mechanical properties of Al4% TiB2 in situ composites synthesized by the optimized fabricate process.
Sample no.

Stirring intensity (rpm)

Stirring start time

Stirring duration (min)

Hardness (HB)

0.2 (MPa)

UTS (MPa)

Elongation (%)

10

180

t0 and t0 45

30

25.17 0.2

72.3 7 1.1

113.6 7 1.4

37.2 7 1.3

Fig. 8. SR-CT images of the Al7% TiB2 composites fabricated: (a) and (b) without stirring; (c) with the developed technique.

particle sedimentation caused by the density difference between


TiB2 (4.5 g/cm3) and Al melt (E 2.7 g/cm3), as a result, TiB2
particles suspend in the melt. During the solidication process,
these suspended particles are likely to be engulfed by the solid
front, nally disperse particles are retained in the solidied matrix.
The mechanical properties of a composite is greatly dependent
on the morphology and size of its second phase particulates.
Considering stress concentration lower mechanical properties,
equiaxed or quasi-equiaxed are most favorable as reinforcing
phases in MMCs. Fig. 9(a) shows that, the particles in the
composite without stirring are mainly in rod-like and plateshape morphology, while rarely in hexagonal. A new TiB2 particle
morphology, cruciform shape, can be detected in the composite.
Both the rod like and cruciform shape conrm particles preferentially grow at specic directions, which is ascribed to the solute
and temperature uctuation exist in the melt. For TiB2 growth,
local enrichment of Ti and B facilitate preferred growth along
/1010S direction [28], which resulted in a rod-like and cruciform
shape morphology. The TiB2 particles formed under the optimized
process condition show highly symmetric hexagonal morphology

(Fig. 9(b)). This equiaxed morphology can be attributed to the


uniform concentration and temperature eld that resulted from
interface stirring, suppress the preferential growth along /1010S
direction in the initial stage of the fabrication process. Although
interface stirring is benecial to particle morphology improvement, no obvious difference in the average particle size can be
found between the stirring and non-stirring composites. Average
particle size for Al4% TiB2 and Al7% TiB2 non-stirring composites
are 270.9 and 316.4 nm, respectively, very similar to those of 261.7
and 327.2 nm in the stirring composites. It should be note in Fig. 9,
no voids or reaction products can be observed near the particle
matrix interface, indicating good interfacial integrity between the
in situ formed TiB2 particles and the aluminum matrix.
3.6. Microstructure and properties of composites with different TiB2
content synthesized by the developed technique
Fig. 10 gives SEM micrographs of two composites with different
particle content (1% and 7%) synthesized by the developed
technique. As can be seen that most of the TiB2 particles distribute

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

365

Fig. 9. SEM micrographs of the TiB2 morphology in situ Al7% TiB2 composites: (a) prepared without and (b) with the developed technique.

Fig. 10. SEM micrographs of the composites: (a) Al1% TiB2; (b) Al7% TiB2.

homogeneously in the aluminum matrix. It is noting that grain


size of the matrix decrease with the increased TiB2 content. As TiB2
content increasing from 0 to 7.0%, the average grain size reduces
from 1077.1 to 78.4 m. The grain renement can be attributed to
the combination of two causes: in the initial state of solidication,
partial of the TiB2 particles act as heterogeneous nucleation
substrate for the aluminum crystals; in the subsequent solidication process, TiB2 particles surround the growing soft aluminum
matrix and act as hard restrictive obstacles against grain growth
[32,33].
The variation of the hardness and tensile properties of the
fabricated AlTiB2 composites as a function of TiB2 content is
depicted in Fig. 11. It is evident that the hardness increases
signicantly with the increasing TiB2 concentration. An improvement of 17%, 37.9% and 63.2% in hardness is observed for 1, 4 and
7% TiB2 reinforced composites respectively, when compared with
the matrix. As it can be seen, the yield strength (0.2) and ultimate
tensile strength (UTS) of the composites are higher than aluminum
matrix, and increase with the increasing amount of TiB2. While the
elongation decreases slightly with the TiB2 content. The 0.2 and
UTS of the composite with Al7% TiB2 are 85.4 MPa and 136.6 MPa,
respectively, which are 260% and 180% higher than those of the
matrix.
The experimental results in terms of percentage change in UTS
and ductility are compared to those reported for TiB2 reinforced
Al-based in situ MMCs available in the literature [17,19,28,3436].
Fig. 12(a) shows that, in spite of lower concentration of reinforcement is incorporated, the AlTiB2 composites in this work show a
higher improvement in strength, indicating that the interface
stirring can offer superior strengthening efciencies to the nal

Fig. 11. Effect of TiB2 content on the mechanical properties of the fabricated
AlTiB2 in situ composites.

composites as compared to the conventional process. Fig. 12(b)


gives a comparison of the elongation decrease among the AlTiB2
composites. It is found that percentage change in elongation is
smaller or comparable in this work with respect to those reported
in the literature. In fact, in present work all the composites
synthesized by the developed interface stirring technique have a
real elongation larger than 28.9%, which is much higher than the
previously reported values.
To investigate the strengthening effects of ne TiB2 particles on
the hardness, tensile properties of the composites, two approaches
have been used to account for the strengthening in the composites
[10]. One is based on the load transferring on particle from matrix
(the continuum mechanics) which result from the good bonding
between the soft matrix and defect free TiB2 particles. The other is
based on the strengthening effect of the particles on the yield

366

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

Fig. 12. Variation of improvement in UTS (a) and the decrease in elongation (b) with TiB2 content for the AlTiB2 in situ composites in this work and in the literature.

commonly be described as [14]




 1=2
 1=2
Hall  Petch k d
 d0

Table 5
Strength calculated basing on each strengthening mechanism.
Composite

Particle
diameter (nm)

Grain size (m)

Hall  Petch

Orowan

CTE

Al1% TiB2
Al4% TiB2
Al7% TiB2

158.1
261.7
327.2

170.7
106.1
78.4

3.4
4.9
6.1

10.3
13.1
15.0

25.4
40.2
48.2

strength of metallic matrix ym (micromechanics strengthening).


Taking into account the load transferring effect of reinforcing
particles, the yield strength of composites (yc) can be expressed
as [2,10]

 

yc ym V p 1 S=2 1  V p 
4
where ym is the yield stress of the reinforced matrix, Vp is the
particle volume fraction, and S is the aspect ratio of the reinforcing
particles, for equiaxed particles S is 1.
Micromechanics strengthening of reinforcement particles on
yield strength of metallic matrix (ym) is due to:
(i) Grain renement strengthening ( Hall  Petch ), grain boundaries can act as obstacles for dislocation slip as well as
dislocation source. TiB2 particles rene the grains of aluminum alloy, thus providing more area to resist the dislocation
movement;
(ii) Orowan strengthening ( Orowan ), large amount of ne TiB2
particles well distribute in the matrix act as obstacles to the
movement of dislocations during the loading of the composites. It is likely for dislocation line to loop around the ne
TiB2 particles by Orowan looping rather than to cut through
them owing to the high elastic modulus (530 GPa) of the TiB2
particles. According to Orowan strengthen mechanism, dislocation loops provide more resistance to the movement of
subsequent dislocations and thus increase the strength of the
composites [32,37];
(iii) Append dislocations ( CTE ) are created due to the difference
of the coefcient of thermal expansion (CTE) between the
matrix and TiB2 particle, these dislocations make the plastic
deformation more difcult.
The yield strength increment caused by crystal grain renment,
Hall  Petch is empirically related to average grain size. Based on
the well-established Hall  Petch relationship, Hall  Petch can

where d and d0 are the grain size of composite and unreinforced


matrix, respectively, k is the HallPetch slope, which is
E74 MPa m1/2 for pure Al. For 7.0% TiB2 reinforced composite,
the calculated strengthening caused by grain renement is
6.1 MPa.
For polycrystalline materials, the increment in yield strength
result from Orowan strengthening can be estimated using the
OrowanAshby equation [38]
Orowan

0:13Gb
D
ln

2b

where G is the matrix shear modulus, b is the Burgers vector, D is


the average diameter of the particles, which increases with the
increasing TiB2 content (Table 5), is the interparticle spacing and
can be expressed as
"
#
1
1 3
1
7
D
2V P
For pure Al, G 26.2 GPa, b 0.286 nm. For Al7.0% TiB2
composite, the calculated yield strength increment caused by
Orowan mechanism is 15.0 MPa.
The strengthening contribution of CTE due to the statistically
stored dislocations introduced by the CTE mismatch between the
matrix and reinforcement can be calculated as [2]
p
CTE Gb
8
Here, is a constant of order 1, is the dislocation density and
can be estimated as [2,38]

12TV P
bD1  V P

where is the difference in CTE between the matrix


(23.6  10  6 K  1) and the reinforcing particles (8  10  6 K  1).
T is the difference between the processing and test temperature.
In this work, the temperatures of liquids Al (660 1C) and the preheated mold (200 1C), are taken as the processing temperature and
the test temperature, respectively.
The calculated strength from each strengthening mechanism is
given in Table 5. It is evident that CTE and Orowan strengthening
contribute mainly the matrix strengthening. While the inuence
from grain renement is not signicant due to the low HallPetch
slope of pure aluminum.
Considering the contribution of several micromechanics
strengthening effects mentioned above, the following equation is

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

367

Fig. 13. SEM fractographs of the aluminum matrix (a), the Al7.0% TiB2 composites produced without stirring (b) and the Al7.0% TiB2 composite fabricated by the developed
technique (c) and (d).

proposed to predict the yield strength of reinforced matrix


q
ym 0 Hall  Petch Orowan 2 CTE 2

10

Where 0 is the yield strength of the unreinforced matrix (sample


0). In present study, measured value of 0 is 24.6 MPa.
After summing the different strengthening results given in
Table 5 using Eq. (10), the calculated yield strength of reinforced
matrix ym is substituted in Eq. (4), then the theoretical yield
strength values of the developed composites are estimated and
plotted in Fig. 9. It is evident that the predicted values estimated by
considering both the load transfer and the contribution from matrix
strengthening are comparable to the experimental values. The
predicted strength value for Al7% TiB2 is 83.0 MPa, which is very
close to the experimental value of 85.4 MPa. Concerning the
contribution of different strengthening mechanisms to the obtained
strength improvements in the nal composites, it is obvious that
CTE is the most effective one, followed by the Orowan strengthening, grain renement strengthening, and load transferring effect.

3.7. Fractograph
Fig. 13 shows the fractographs of the unreinforced matrix,
Al7% TiB2 composites produced with and without the developed
stirring technique. As shown in Fig. 13(a), unequally distributed
large and deep equiaxed dimples dominate in the fracture of the
matrix, indicating a large degree of plastic deformation. During
the tensile test, microvoids initiate and grow with increasing load,
when these voids reach a critical size to coalesce, the nal fracture
takes place and dimples are retained in the fracture surface. Thus,

the degree of plastic deformation and the ductility of the matrix


depend on the extent of the growth of microvoids.
Shown in Fig. 13(b), agglomerations, dimples as well as cleavage
facets are manifested in fracture surface of non-stirring composite,
which has the low UTS of 90.8 MPa and poor elongation of 16.4%.
Cracks can be observed at the matrixagglomerate interface, suggesting that cracks initiate around the TiB2 agglomerates and then
propagate along the matrixagglomerate boundary, nally leading
to a brittle fracture. Gathered at cleavage facets are also found, EDS
analysis reveals these facets formed due to brittle rupture of coarse
TiB2.
In the case of the MMCs synthesized with the new technique,
large quantities equiaxed dimples surrounded by tiny dimples
dominate in the fracture surface (Fig. 13(c)), indicating that the
composites fracture still follow the ductility of the aluminum matrix
despite the stiff reinforcement is incorporated. It should be note
these equiaxed dimples are much shallower and smaller as compared to those in non-reinforced matrix. This can be attributed to
TiB2 particles act as effective barriers to the growth and coalescence
of the voids. High magnication image (Fig. 13(d)) shows that tiny
dimples enclosing particles deform parallel to sliding direction,
which could be a result of the severe deformation of the matrix to
maintain interface integrity. All these results mentioned above
provide rm evidence that, the excellent wetting and strong
bonding between the matrix and reinforcements can be achieved
using the developed salts/aluminum interface stirring technique.

4. Conclusions
(1) Processing parameters of salts/aluminum interface mechanical
stirring have substantial effects on microstructure and

368

F. Chen et al. / Materials Science & Engineering A 625 (2015) 357368

mechanical properties of the nal composites. Proper stirring


intensity and duration are helpful to suppress the formation of
severe agglomeration, meanwhile, to prevent inclusion
entrapping and melt oxidation. TiB2 particle morphology is
mainly determined in the rst 15 min, while particle distribution affected by the whole holding process.
(2) Interface mechanical stirring signicantly improves TiB2 particles distribution in aluminum matrix, defect-free TiB2 reinforced in situ composites can be produced by applying stirring
with speed of 180 rpm in both the rst and the last 15 min of
the 60 min composite synthesis process.
(3) SR-CT is a powerful tool in the investigation of stereoscopic
structure of in situ TiB2 reinforced Al based composites, a
much better vision of TiB2 distribution has been attained.
(4) The fabricated composites possess much higher improvements
in mechanical properties when compared with the aluminum
matrix and the AlTiB2 in situ composites fabricated by
conventional process. The enhancement in the mechanical
properties mainly result from CTE , followed by Orowan
strengthening, grain renement strengthening and load transferring effect.

Acknowledgments
This work was supported by the National Natural Science
Foundation of China (Nos. 51071035, 51274054, 51375070 and
U1332115), and the Key grant Project of Chinese Ministry of
Education (No. 313011). It is a pleasure to thank Jingyu Han for
his help in the SEM analysis, and Guangyuan Yan for his help in
preparing the SR-CT specimens of the work. The authors also wish
to thank the Shanghai Synchrotron Radiation Facility for provision
of synchrotron radiation facilities. Special thanks to all the staff
members of the BL13W1 beamline of SSRF for their contribution
during the synchrotron experiments.
References
[1] Y. Zhang, N. Ma, H. Wang, Y. Le, X. Li, Mater. Des. 28 (2007) 628632.
[2] S. Jayalakshmi, S. Gupta, S. Sankaranarayanan, S. Sahu, M. Gupta, Mater. Sci.
Eng. A 581 (2013) 119127.

[3] S.A. Sajjadi, H.R. Ezatpour, M. Torabi Parizi, Mater. Des. 34 (2012) 106111.
[4] P. Poddar, V.C. Srivastava, P.K. De, K.L. Sahoo, Mater. Sci. Eng. A 460461 (2007)
357364.
[5] S. Gopalakrishnan, N. Murugan, Compos. Part. B-Eng. 43 (2012) 302308.
[6] S. Kumar, M. Chakraborty, V. Subramanya Sarma, B.S. Murty, Wear 265 (2008)
134142.
[7] C.S. Ramesh, A. Ahamed, Wear 271 (2011) 19281939.
[8] M. Emamy, M. Mahta, J. Rasizadeh, Compos. Sci. Technol. 66 (2006)
10631066.
[9] S. Lakshmi, L. Lu, M. Gupta, J. Mater. Process. Tech. 73 (1998) 160166.
[10] Q. Zhang, B.L. Xiao, W.G. Wang, Z.Y. Ma, Acta Mater. 60 (2012) 70907103.
[11] D.G. Zhao, X.F. Liu, Y.C. Pan, X.F. Bian, X.J. Liu, J. Mater. Process. Tech. 189 (2007)
237241.
[12] S. Suresh, N. Shenbag, V. Moorthi, Procedia Eng. 38 (2012) 8997.
[13] Y. Zhang, N. Ma, H. Wang, Mater. Lett. 61 (2007) 32733275.
[14] T. Wang, Z. Chen, Y. Zheng, Y. Zhao, H. Kang, L. Gao, Mater. Sci. Eng. A 605
(2014) 2232.
[15] A. Mandal, R. Maiti, M. Chakraborty, B.S. Murty, Mater. Sci. Eng. A 386 (2004)
296300.
[16] D. Zhao, X. Liu, Y. Liu, X. Bian, J. Mater. Sci. 40 (2005) 43654368.
[17] J. Xue, J. Wang, Y. Han, P. Li, B. Sun, J. Alloys Compd. 509 (2011) 15731578.
[18] F. Wang, J. Xu, J. Li, X. Li, H. Wang, Mater. Des. 33 (2012) 236241.
[19] K.L. Tee, L. L, M.O. Lai, Mater. Sci. Eng. A 339 (2003) 227231.
[20] M. Habibnejad-Korayem, R. Mahmudi, W.J. Poole, Mater. Sci. Eng. A 519 (2009)
198203.
[21] S. Amirkhanlou, B. Niroumand, Trans. Nonferr. Met. Soc. China 20 (2010)
s788s793.
[22] F. Xu, Y. Li, X. Hu, Y. Niu, J. Zhao, Z. Zhang, Mater. Lett. 67 (2012) 162164.
[23] M.Y. Wang, Y.J. Xu, T. Jing, G.Y. Peng, Y.N. Fu, N. Chawla, Scr. Mater. 67 (2012)
629632.
[24] R.W. Hamilton, M.F. Forster, R.J. Dashwood, P.D. Lee, Scr. Mater. 46 (2002)
2529.
[25] J. Baruchel, J.Y. Bufere, P. Cloetens, M. Di Michiel, E. Ferrie, W. Ludwig,
E. Maire, L. Salvo, Scr. Mater. 55 (2006) 4146.
[26] L. Salvo, P. Cloetens, E. Maire, S. Zabler, J.J. Blandin, J.Y. Bufere, W. Ludwig,
E. Boller, D. Bellet, C. Josserond, Nucl. Instrum. Meth. B 200 (2003) 273286.
[27] Z. Sadeghian, M.H. Enayati, P. Beiss, J. Mater. Sci. 44 (2009) 25662572.
[28] Z. Chen, T. Wang, Y. Zheng, Y. Zhao, H. Kang, L. Gao, Mater. Sci. Eng. A 605
(2014) 301309.
[29] J. Fjellstedt, A.E.W. Jarfors, Mater. Sci. Eng. A 413 (2005) 527532.
[30] C.F. Feng, L. Froyen, J. Mater. Sci. 35 (2000) 837850.
[31] E.-M. Nahed, M.A. Taha, A.E.W. Jarforsb, H. Fredriksson, J. Alloys Compd. 292
(1999) 221229.
[32] L. Lu, M. LA1, F.L. CHEN, Acta Mater. 45 (1997) 42974309.
[33] H.B. Michael Rajan, S. Ramabalan, I. Dinaharan, S.J. Vijay, Mater. Des. 44 (2013)
438445.
[34] K.L. Tee, L. Lu, M.O. Lai, J. Mater. Process. Tech. 90 (1999) 513519.
[35] A.R. Kennedy, A.E. Karantzalis, S.M. Wyatt, J. Mater. Sci. 34 (1999) 933940.
[36] S.C. Tjong, K.F. Tam, Mater. Chem. Phys. 97 (2006) 9197.
[37] J. Lee, J.Y. Jung, E.-S. Lee, W.J. Park, S. Ahn, N.J. Kim, Mater. Sci. Eng. A 277
(2000) 274283.
[38] Z. Zhang, D. Chen, Scr. Mater. 54 (2006) 13211326.

You might also like