You are on page 1of 9

Int.

Journal of Refractory Metals and Hard Materials 43 (2014) 250–258

Contents lists available at ScienceDirect

Int. Journal of Refractory Metals and Hard Materials


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

Spark plasma sintering of (Ti0.9W0.1) C powders prepared by


mechanical alloying
C. Slama a,b,⁎, M.D. Boulaares a, A. Teber a, F. Schoenstein c, M. Abdellaoui a, N. Jouini c
a
Laboratoire des matériaux utiles, Institut National de Recherche et d’Analyse Physico-chimique, Pôle technologique de Sidi Thabet, 2020 Sidi Thabet, Tunisia
b
Ecole Supérieure des Ingénieurs de l’Equipement Rural, Medjez El Bab, 9070, Tunisia
c
Laboratoire des Sciences des Procédés et des Matériaux, CNRS-UPR 9001, Université Paris 13, 93430 Villetaneuse, France

a r t i c l e i n f o a b s t r a c t

Article history: Nanocrystalline (Ti0.9W0.1)C powder with a diffraction crystallite size of about 10 nm was synthesized by
Received 20 September 2013 mechanical alloying. The formation of (Ti0.9W0.1)C carbide was detected by XRD measurements and microscopic
Accepted 17 December 2013 observation. The sintering of these powders by a spark plasma sintering (SPS) at different temperatures were also
Available online 24 December 2013
studied. The results show that the maximum hardness was obtained for more relative density materials,
meanwhile, the grain size is large. The micro-hardness and the relative density of the powder milled for 10 h
Keywords:
Ti0.9W0.1C carbide
and sintered at 1200 °C for 5 min under 100 MPa reach, respectively, 2978 HV and 98.35%.
Mechanical alloying © 2013 Elsevier Ltd. All rights reserved.
Spark plasma sintering
Density
Hardness

1. Introduction synthesizing nanocrystalline carbide materials at room temperature


[9,10].
Titanium carbide is an excellent material for cutting tool applications: Many attempts have been made to synthesize (Ti,W)C powders via
it has a high hardness (Vickers Hardness, HV = 3200 kgf/mm2), high the thermal reaction of TiC/WC powder mixture at high temperatures
melting temperature (3260 °C), low density, high thermal and or by the self heat-propagation synthesis (SHS) of a Ti, W and C powder
electrical conductivity, high chemical and thermal stability, high wear mixture [11,12]. Since TiC is thermodynamically more stable than WC,
resistance and high solubility compared with other carbides [1]. (Ti,W)C is synthesized commercially by reacting TiC with WC in the
Owing to this excellent combination of properties, TiC is often used in temperature range of 1700–2100 °C for maintain duration ranging
abrasives, high-speed cutting tools, electrical discharge machining, from 10 to 20 h [11]. Jung and Kang [6] synthesized (Ti,W)C powders
grinding wheels and coated cutting tips [2,3]. The applications are, by milling a mixture of carbon, TiO2 and WO3 and reducing at 1200 °C
however, significantly limited by its intrinsic brittleness [4]. Therefore, for 1 h. Then, the powder was consolidated into pellets, using a cold-
some other elements are added to improve the ductility of TiC. Most press at pressure of 125 MPa and sintered at 1510 °C for 1 h under sec-
used elements for substitution are tungsten and nitrogen. Tungsten ondary vacuum (10−4 atm). However, the synthesis process of all these
was tested as a dopant, yielding Ti1 − xWxC [5,6]. Nitrogen was mentioned techniques was either time consuming or involving high
introduced into the system to inhibit the growth of carbide particles, pressure and high temperature, and always produces inhomogeneous
yielding Ti1 − xWxC1 − y Ny [7,8]. Due to high melting temperatures of powders. In addition, no attempts on direct densification of nanocrystal-
α-Ti (1668 °C), W (3407 °C) and graphite (3826 °C), TiC and (Ti,W)C line (Ti,W)C particles were reported.
are hardly prepared by conventional metallurgical methods and they In the other hand, Spark Plasma Sintering (SPS) is a fast powder
require expensive high temperature equipment. However, mechanical consolidation technique where powders are simultaneously heated
alloying (MA) is relatively a new technique and being utilized success- and compacted by uniaxial pressing [13,14]. The heating results from
fully in preparing materials which are very difficult to produce by any a pulsed electric field and a high heating rate of more than 100 K/min
conventional method due to high melting temperatures of elements. can be achieved. This technique offers the possibility to produce bulk
Moreover, it is considered to be the most powerful technique for bodies with high density at lower sintering temperatures and with
shorter sintering times.
The objectives of the present work are (1) the production of
nanocrystalline Ti0.9W0.1C carbide by high-energy ball milling the
⁎ Corresponding author at: Laboratoire des matériaux utiles, Institut National de
Recherche et d’Analyse Physico-chimique, Pôle technologique de Sidi Thabet, 2020 Sidi
elemental α-Ti, W and C (graphite) powders at room temperature, (2)
Thabet, Tunisia. Tel.: +216 98900018. the characterization of the microstructure of the prepared materials and
E-mail address: chihebslama@yahoo.fr (C. Slama). (3) finally the study of the mechanical alloying (MA) and temperature

0263-4368/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijrmhm.2013.12.009
C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258 251

effects on the Ti0.9W0.1C carbide formation by spark plasma sintering


from Ti, W and C powders.

2. Experimental methods

Elemental powder mixtures of Ti (b40 μm, 99.9%, Prolabo), W


(1–5 μm, 99.9%, Metal basis) and carbon (99.9%, Fischer scientific)
were sealed into a stainless steel vial (45 ml in volume) with 5 stainless
steel balls (15 mm in diameter and 14 g in mass) in a glove box filled
with purified argon. The ball to powder weight ratio was about 70 to
1. The mechanical alloying (MA) procedure was performed at room
temperature using a planetary ball mill (Fritsch pulverisette 7) with
milling conditions corresponding to 7.41 10− 2 J/hit kinetic shock
energy, 70.1 Hz shock frequency and 5.19 W/g injected shock power
[15,16]. The obtained nanocrystalline Ti0.9W0.1C carbide powders were
sintered using (SPS)—5155 SYNTEX apparatus at 1150, 1200 and
1250 °C for 5 min under a pressure value of 100 MPa.
The crystalline properties of the powders and the bulk samples were
characterized by X-ray diffraction using a ( –2 ) Panalytical XPERT PRO
MPD diffractometer operating with Cu Kα radiation (λ = 0.15406 nm).
The existing phases were identified by High Score Plus program based
on the ICDD PDF2 data base. The diffraction crystallite size was
determined by the FullProf program [17] using the Rietveld method
[18]. The morphology was studied by scanning electron microscopy
using FEI Quanta 200 environmental scanning electron microscope
Fig. 1. XRD patterns of mechanically alloyed Ti0.9W0.1C powders after various MA
coupled with EDAX. The microstructure of the alloys was studied by FEI durations.
Tecnai G2 high resolution transmission electron microscope (HRTEM)
operating at 200 KV. Specimens were prepared by crushing the samples
under alcohol. A drop of the aqueous suspension was put on a holey 3. Results and discussion
carbon film supported by a copper grid.
The relative densities of the consolidated samples were determined 3.1. Synthesis and characterization of nanocrystalline Ti0.9W0.1C carbide by
by the Archimedes method using O-xylene as measuring liquid (density mechanical alloying
of 0.88 at 25 °C).
Microhardness measurements were conducted using a Duramin 20 The XRD patterns of mechanically alloyed Ti0.9W0.1C alloy powders
Vickers device under a test force of 2.94 N (HV 0.3) during 7 s. Ten obtained after selected MA duration are presented in Fig. 1. The XRD
measurements were taken for each reported Vickers hardness value. pattern of the starting reactant materials (Δt = 0 h), used as a

Table 1
Final refinement results of Ti0.9W0.1C.

Mechanical alloying duration (h) Phase present Space group Lattice parameter (nm) Weight (%) DCS (nm) Reliability factors χ2

1.5 (Ti,W)C Fm-3 m a = 0.43218(1) 93.08 ± 0.75 49 RB = 3.09 10.1


Rf = 1.22
W2C P-31 m a = 0.52045(6) 2.23 ± 0.08 33 RB = 4.05
c = 0.4736(1) Rf = 2.41
W Im-3 m a = 0.31681(1) 4.69 ± 0.10 34 RB = 0.626
Rf = 0.417
2.5 (Ti,W)C Fm-3 m a = 0.43212(4) 97.22 ± 2.36 16 RB = 1.50 6.04
Rf = 0.767
W Im-3 m a = 0.31688(4) 2.78 ± 0.26 25 RB = 4.27
Rf = 2.22
4 (Ti,W)C Fm-3 m a = 0.43180(5) 95.86 ± 2.33 13 RB = 0.391 1.58
Rf = 0.192
W Im-3 m a = 0.31668(5) 2.46 ± 0.22 20 RB = 4.15
Rf = 1.48
Fe Im-3 m a = 0.28737(9) 1.68 ± 0.18 19 RB = 0.123
Rf = 0.061
7 (Ti,W)C Fm-3 m a = 0.43168(3) 92.34 ± 1.94 12 RB = 0.687 2.25
Rf = 0.354
W Im-3 m a = 0.31668(4) 1.99 ± 0.31 17 RB = 1.10
Rf = 0.698
Fe Im-3 m a = 0.2872(5) 5.66 ± 1.00 16 RB = 1.10
Rf = 0.736
10 (Ti,W)C Fm-3 m a = 0.43034(2) 91.63 ± 0.78 8 RB = 0.504 1.31
Rf = 0.242
Fe Im-3 m a = 0.2879(3) 8.37 ± 0.32 14 RB = 0.370
Rf = 0.292
20 (Ti,W)C Fm-3 m a = 0.42947(2) 91.31 ± 0.81 6 RB = 0.740
Rf = 0.3601.15
Fe Im-3 m a = 0.2877(5) 8.69 ± 0.36 11 RB = 1.52
Rf = 0.639
252 C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258

reference, consists of Bragg diffraction peaks of Ti, C and W. After 30 min


of MA, all the graphite Bragg lines disappeared except the (002) plane
situated at 26.67° in 2 . After 1.5 h of MA time, the graphite and Ti
Bragg lines disappeared totally, only W peaks remained and W2C
carbide appeared. At the same time, a new phase with a face center
cubic NaCl type structure and corresponding to the (Ti,W)C carbide
has appeared. After milling for 2.5 h, W peaks remained, W2C peaks
disappeared completely and (Ti,W)C peaks broadened due to the
refinement of the grains and the introduction of internal strain. After
milling for 4 h, only the (TiW)C and W peaks exist with the appearance
of new peaks related to austenite Fe as contamination resulted from the
wear between the balls and the vial surfaces. The increasing of the
milling time up to 10 h led to the disappearance of W peaks. Further
milling (20 h) induced a decrease in the diffraction crystallite size of
the Ti0.9W0.1C carbide indicated by the very broad Bragg peaks. Thus,
the sequence of phase formation with milling time for this W substituted
carbide is:

1:5 h 2:5 h 4h
0:9Ti þ 0:1 W þ C → W2 C þ W þ ðTi; WÞC → W þ ðTi; WÞC → W þ ðTi; WÞC þ Fe
10 h
→ Ti0:9 W0:1 C þ Fe:

Table 1 summarizes the characteristic parameters of the final


refinement results of Ti0.9W0.1C (DCS: diffraction crystallite size, RB:
Bragg R–factor, Rf: Rf–factor, χ2: Chi2).
Fig. 2 shows the Rietveld refinement of the Ti0.9W0.1C powders MA
for a) 10 h and b) 20 h. The upper continuous line is the calculated
profile and the lower continuous line is the difference between the
experimental and the calculated profiles. Rietveld refinements (Fig. 2)
clearly show that W substitutes Ti in the TiC compound. As a
consequence, the lattice parameter of the (Ti,W)C carbide would be
expected to decrease, because W has smaller atomic radius (0.137 nm)
than that of the Ti atom (0.145 nm).
Fig. 3 shows the variation of the diffraction crystallite size (DCS) of
the Ti0.9W0.1C carbide and the weight content of the different phases
against the MA duration. As shown in Table 1, when increasing
the milling duration from 1.5 up to 10 h, the lattice parameter of
Ti0.9W0.1C decreases from 0.43218 nm to 0.43034 nm and the DCS
decreases from 49 to 8 nm (Fig. 3). So, the lattice parameter decrease Fig. 2. Example of FullProf program fit of the Ti0.9W0.1C powders MA for a) 10 h and b)
20 h.
of Ti0.9W0.1C is related to the substitution of Ti by W and consequently
to the formation of Ti0.9W0.1C carbide, whereas the decrease of (DCS)
can be attributed to the fracture of starting powder particles. milling
for 10 h, the lattice parameter of Ti0.9W0.1C remain constant while the
grain size decreases very slowly from 8 (10 h) to 6 nm (20 h). It is
suggested that an equilibrium was reached between the welding and
the fracturing phenomenon after milling for 10 h. According to Ghosh
and Pradhan [19], the decrease of lattice parameter of TiC after milling
may be attributed to the compressive stress generated in MA process.
Rahaei et al. [20] reported that the abrupt decrease in particle size of
TiC can be explained by the fracture of TiC particles during the milling
process and thus due to the brittleness behavior of TiC carbide.
As shown in Fig. 3, when increasing the milling duration from 1.5 up
to 2.5 h, the weight content of W2C phase decreases from 2.23 wt.% to
zero and the weight content of W decreases from 4.69 to 2.78 wt.%
in favor of the increase of the formation of the Ti0.9W0.1C carbide
(97 wt.%). Further milling up to 10 h results in full formation of
Ti0.9W0.1C carbide with weight content value equals to about 92%.
After milling for 4 h, we note a contamination with iron. So, when
increasing the milling duration from 4 h up to 20 h, the weight content
of Fe increases from 1.68% to 8.69%.
Fig. 4 shows SEM micrographs in secondary electron mode of
Ti0.9W0.1C carbide obtained after 10 h and 20 h of milling. For milling
duration at 10 h, MA induces the formation of large agglomerates com-
posed by fine particles (Fig. 4a) as a result of repeated welding,
fracturing and rewelding operations. By increasing the milling time up Fig. 3. Variation of the diffraction crystallite size of the Ti0.9W 0.1C powders and the weight
to 20 h, the morphology of powders is still homogenous but with very content of different phases against the milling time.
C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258 253

Fig. 4. SEM micrographs of Ti0.9W0.1C obtained after a) 10 h and b) 20 h. c) EDX of the powder Ti0.9W0.1C obtained after 10 h.

fine agglomerated particles smaller than 1 μm in size (Fig. 4b). The This TEM result (5–10 nm) confirms those obtained by XRD measure-
corresponding energy dispersive X-ray microanalysis (EDX) of the ment (6–8 nm).
Ti0.9W0.1C powder obtained after 10 h is shown in Fig. 4c. The X-ray
microanalysis shows the existence of the major peaks corresponding 3.2. Consolidation and microstructure of bulk samples
to Ti, W and C. We also note the existence of peaks of Fe and Cr due to
the wear between the balls and the vial surfaces. The chemical analysis Table 3 summarizes the SPS process conditions and the values of the
results are presented in Table 2. The calculated mean composition of the obtained physical properties (density of compacted powders also called
carbide is Ti = 42.96 at.%, W = 4.87 at.% and C = 52.16 at.% which experimental density, relative density and diffraction crystallite size).
correspond to a mean formula of (Ti0.898W0.101)C1.09. The theoretical density of Ti0.9W0.1C by the following equation:
Figs. 5 and 6 indicate TEM micrographs of Ti0.9W0.1C powder
obtained after milling for respectively 10 and 20 h. The bright-field n ½0:9M Ti þ 0:1MW þ MC 
micrograph (Fig. 5) shows that agglomerated powders exhibit different ρthe ¼
NV
particle sizes with an irregular shape of sample obtained after 10 h.
Bright-field micrograph of Ti0.9W0.1C powder obtained after 20 h of where n = number of atoms per unit cell; MTi, MW and MC = the
milling (Fig. 6) shows many agglomerated powders with an irregular atomic weight of titanium, tungsten and carbon, respectively; V = the
shape having different particles with an average size about 5 to 10 nm. cell volume of Ti0.9W0.1C; and N = Avogadro's number. The relative
density (ρrel) of the sample can be calculated by the equation:
Table 2
The elementary constituents of the (TiW)C phase determined by EDX.
ρrel ¼ ρexp =ρthe :
Elements C Ti W Fe Cr

Weight percentage (%) 15.74 51.69 22.47 8.76 1.34 The diffraction crystallite sizes (d) of the sintered materials were
Atomic percentage (%) 48.63 40.05 4.54 5.82 0.95
determined by the FullProf program using the Rietveld refinement.
254 C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258

Table 3
Experimental conditions adopted during the SPS process and corresponding values
of physical properties (MAd, mechanical alloying duration; T, sintering temperature; t,
holding time; P, pressure; ρ, density of the compacted sample; ρrel, corresponding relative
density; d, diffraction crystallite size obtained by Rietveld refinement).

MAd T t P ρ ρrel d
−3
(h) (°C) (min) (MPa) (g cm ) (%) (nm)

10 1200 5 100 6.02 98.35 38.14


20 1150 5 100 5.75 93.42 20.18
20 1200 5 100 5.80 94.31 23.12
20 1250 5 100 5.83 94.68 25.50

densification started around 900 °C. Indeed, the displacement increases


very quickly from 0.35 mm to about 1.88 mm when increasing the
temperature from 900 °C to 1200 °C (750 s). This displacement will
be attributed first to the default restoration which leads to the material
densification. This is facilitated by the thermal softening which leads to
a material with smooth and round shape grain. For holding time ranging
from 750 to 1030 s, the displacement remains constant as the material
reaches its maximum density. Further displacement occurred during
the cooling down of the SPS temperature. According to Chaim et al.
[22], the observed increase of the displacement in the end of the process
can be related to the sum of the thermal shrinkages of the dense
specimen and the graphite assembly.
Fig. 8 shows the displacement profile recorded during the SPS
consolidation of Ti0.9W0.1C nano-powders MA for 20 h and sintered at
1200 °C for 5 min under 100 MPa. During the SPS process, the pressure
Fig. 5. Bright-field micrograph of milled powders obtained after 10 h.
was set to 50 MPa during the first 170 s. Then, it was increased to
100 MPa and remained constant between 360 and 1570 s (Fig. 8a).
Fig. 7 shows the displacement profile recorded during the SPS
consolidation of Ti0.9W0.1C nano-powders MA for 10 h and sintered at
1200 °C for 5 min under 100 MPa. During the SPS process, the pressure
was set to 50 MPa and was maintained until 190 s (Fig. 7a). Then, it was
increased to 100 MPa and maintained constant from 360 s to 1570 s.
The temperature was maintained constant at 600 °C during the first
360 s. Then, it was increased to 1200 °C and was remained constant
from 750 s to 1030 s. As it can be seen during the first 360 s, the
displacement increases to 0.35 mm when increasing the pressure
from 50 to 100 MPa, whereas the temperature remains constant
(600 °C). As the pressure increases and the temperature remains
unchanged, we assume that this displacement is attributed to cold
compaction of the initial powder. Similar results have been observed
by Abderrazak et al. [21] and Teber et al. [10]. Then, the displacement
remains constant until 530 s (0.35 mm). Fig. 7b shows that the

Fig. 7. Variation of a) the pressure and the displacement and b) the temperature and the
displacement, as a function of holding time during SPS of Ti0.9W0.1C nano-powders MA
Fig. 6. Bright-field micrograph of milled powders obtained after 20 h. for 10 h.
C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258 255

Fig. 9 shows a superposition of the XRD patterns of Ti0.9W0.1C


powders MA for 10 h and 20 h after sintering at various temperatures.
For samples milled for 10 h and 20 h after sintering at respectively
1200 and 1150 °C, no structural changes were observed. On the contrary,
for samples milled for 20 h a thermal decomposition is observed after
sintering at 1200 and 1250 °C. So, a part of W substituting Ti in (Ti,W)C
carbide dissolves. According to Teber et al. [24,25], for a (Ti1 − x Zrx) C
powders sintered at 1650 °C, a thermal decomposition of Zr is observed
for x = 0.25. Similarly, the solubility limit, x, of W in the (Ti1 − xWx)C
phase is reported to be ~0.5 [6]. When it exceeds this limit, pure WC
co-exists with (Ti,W)C in the microstructures.
The precipitation of elemental crystalline W can be attributed to the
excess of cumulated energy resulting first from the non consumed
shock energy and second from the exothermic reaction heat flow.
Jung and Kang [6] considered that high-energy milling transfers the
energy to the powder mixture, resulting in the formation of metastable
phases. In addition, according to Shatov et al. [26], (Ti,W)C formation is
a typical exothermic reaction with a large negative enthalpy estimated
for
to about ΔH = −274 kJ mol−1.
ðTiW ÞC
In order to interpret these results, one can explore the effects of
ball-milling on the formation mechanism of (Ti,W)C carbides and on
the SPS temperatures. At 20 h of MA time, the milling causes a high defect
density (i.e. dislocations, stacking faults, grains boundaries, etc.,…) inside
the powder particles, which accelerate the interdiffusion of all elements
during sintering and promote the particle refinement of the formed
phases. At 10 h of MA, all the W diffuses into the cell to form stable
compound. So, the W does not appear with annealing at 1200 °C. At
20 h, the mechanical alloying with high energy promotes the location
of W in the grain boundaries in amorphous form. This amorphous W
crystallizes with annealing at 1200 °C. Ohser-Wiedemann et al. [27]
Fig. 8. Variation of a) the pressure and the displacement and b) the temperature and the synthesized Mo-W powders by mechanical alloying and have reported
displacement, as a function of holding time during SPS of Ti0.9W0.1C nano-powders MA
for 20 h.
that when the milling time increased from 0 h to 40 h, the dislocation
densities increased up to 1012 cm−2, and simultaneously, the crystallite
size decreased down to 20 nm. In the present work, after 20 h milling,
The temperature was maintained constant around 600 °C until 360 s the diffraction crystallite size decreases down to 6 nm as a result of a
and then increased to 1200 °C and remained constant from 750 s to high dislocation densities. Similar phenomenon has also been reported
1030 s (Fig. 8b). During the first 250 s, the displacement slightly by Ohser–Wiedemann et al. [28] concerning (Ti1 − xMox) C carbides.
increased to 0.15 mm under the pressure (66 MPa) effect. The The morphology of the consolidated binary carbide (Ti0.9W0.1) C
temperature was kept constant to 600 °C (the temperature was too sintered by SPS from nanocrystalline alloys obtained after 10 h and
low to induce any phase transformation). So, this displacement is 20 h MA is shown in Fig. 10. The consolidated carbide obtained from
probably attributed to the nano-particle rearrangement. Then, the powders milled for 20 h and sintered at 1150 °C (Fig. 10a) presented
displacement increases from 0.15 mm to about 0.42 mm when
increasing the pressure from 66 MPa (600 °C) to 100 MPa (620 °C).
We assume that in this stage the displacement value is affected only
by the pressure increase as the temperature increases only from 600
to 620 °C. In the stage 400–750 s, the displacement raised up from
0.42 mm to 1.65 mm and two densification phenomena can be
discerned. First, at 680 °C (400 s), the temperature is sufficiently high
to allow the default restoration and the density increase leading to a
rapid displacement increase from 0.42 mm to about 0.85 mm when
the temperature increases from 680 °C (400 s) to 1000 °C (605 s).
Between 1000 °C (605 s) and 1200 °C (750 s), a rapid displacement
increase of 1.65 mm is observed. This quick increase will be attributed
to a progressive densification from a dense material to a more dense
one. Finally, the displacement slightly increased to 1.70 mm when the
temperature increases to 1200 °C which is probably attributed to the
densification of the formed phases. As reported above, the observed
increase of the displacement in the end of the process is due to the
thermal shrinkages of the dense specimen and the graphite assembly.
Figs. 7 and 8 show a decrease of the onset temperature of densifica-
tion with the increase of the milling time. In fact, the onset of the
densification temperature decreases from 900 °C for 10 h of milling to
about 680 °C for 20 h of milling. This decrease of sintering temperature
is due to the smallest grain size and the most uniform microstructure. Fig. 9. XRD patterns of Ti0.9W0.1C MA for 10 h and 20 h and sintered at different
Similar phenomenon has also been reported by Han et al. [23]. temperatures during 5 min.
256 C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258

Fig. 10. SEM micrographs of Ti0.9W0.1C sintered by SPS at different temperatures from materials obtained by MA after 20 h at a) 1150 °C, b) 1200 °C, c) 1250 °C and after 10 h at d) 1200 °C.

small pores (~1.5 μm) with large density. For sintering temperatures of Fig. 11 shows the variation of diffraction crystallite size and relative
1200 °C and 1250 °C (Fig. 10b and c), the small pores disappeared density of the MA carbides as a function of sintering temperature under
progressively and large ones (~ 2–2.5 μm) with low density were 100 MPa and 5 min. The crystallite size and the relative density of the
observed. The pore size and density were reduced (~ 1 μm) for (Ti,W)C carbide obtained by SPS at 1200 °C after 10 h of MA are equal
consolidated carbide obtained from powders milled during 10 h and respectively to 38.14 nm and 98.35%. As milling time increases (20 h),
sintered at 1200 °C (Fig. 10d) as compared for those obtained for 20 h. the crystallite size of the (Ti,W)C carbide formed during SPS are lesser
as compared to 10 h and increases slightly from 20.18 nm to 25.5 nm
when the sintering temperature increases from 1150 to 1250 °C. As a
consequence, the relative density increases progressively from 93.42%
to 94.68%. Based on these results, it may be seen that for the same
sintering temperature (1200 °C) as milling time increases, the crystallite
size and subsequently the density decrease. Also, for the same alloying
duration, as the sintering temperature increases, the crystallite size and
consequently the relative density increase.
Teber et al. [10] considered that diffusion bonding at the particle
boundaries and other diffusivity paths are enhanced at higher tempera-
tures, which could affect densification and grain growth of TiC ceramics.
According to Locci et al. [29] for a composite TiC–TiB2, as milling time
increases, interfacial area between reactants increases and powder size
decreases. These phenomena induced by mechanical treatment enhance
sintering processes and, consequently material densification.

3.3. Mechanical properties of bulk materials

Fig. 12(a) shows the variation of the Vickers micro-hardness of


Ti0.9W0.1C carbide as function of SPS temperature. The Vickers micro-
hardness (HV) of Ti0.9W0.1C powders milled for 10 h is equal to
2978 HV for SPS temperature equal to 1200 °C. For (Ti,W)C obtained
after 20 h of milling, the HV value increases progressively from
2745 HV to 2871 HV when increasing the temperature from 1150 to
Fig. 11. Variation of the diffraction crystallite size and the relative density of sintered 1250 °C. So, for SPS at 1200 °C, the increase of the MA duration from
Ti0.9W0.1C powders as a function of sintering temperatures. 10 to 20 h induces the HV decrease from 2978 HV to 2846 HV. Teber
C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258 257

our results show that the hardness indentations are about 13–14 μm,
whereas the pores size is about 1–3 μm (Fig. 10), as reported above.
In order to further investigate the influence of sintering temperature
on crystallite size of (Ti0.9W0.1) C phase, the lattice parameter of
(Ti0.9W0.1) C phase with different temperature was experimentally
calculated from the XRD analysis. The values of lattice parameter
calculated before and after SPS are listed in Table 4.
As shown in Table 4, the lattice parameter values slightly increase
with sintering temperature in comparison to those obtained before
sintering, as a result of the residual strain release. However, as the
sintering temperature increases, the crystallite size increases, the pore
size and density decrease leading to an improvement of the relative
density.
According to S. Park et al. [30], TiC is a nonstoichiometric compound,
and fully annealed TiC shows a Ti:C mole ratio of 1:0.98. The carbon
deficiency in the compound varies with temperature and the addition
of other elements in the form of a solid solution. In general, the carbon
content decreases with increasing W content in (Ti,W)C. Generally,
the addition of W is reported to reduce the lattice parameter in the
region of (Ti,W)C formation [31]. Both the small size of the W atom in
the TiC structure and the loss of carbon cause the decrease in the lattice
parameter. The lattice parameter of annealed TiC is 0.43274 nm [32]. In
this study, the lattice parameters of the (Ti,W)C sintered are larger than
those of the (Ti,W)C powders with the same composition. In addition, as
mentioned above in (EDX) results, the composition of (Ti,W)C powder
has the following formula: (Ti0.898W0.101)C1.09. This formula shows
clearly a carbon excess which explain the increase of the lattice
parameter. Biedunkiewicz and Wrobel [33] and Bodziony et al. [34]
found that the growing grain size is accompanied by increasing lattice
constants. As shown in Table 3, the grain size, in our present work,
increases with sintering temperature. Therefore, this result is in good
agreement with Biedunkiewicz and Wrobel [33] and Bodziony et al.
[34] investigations. So, the lattice parameter increase of Ti0.9W0.1C
carbide is affected by the sintering temperature and by the excess
Fig. 12. (a) Variation of Vickers micro-hardness of sintered Ti0.9W0.1C powders versus carbon content. However, after 10 h of milling, the lattice parameter
sintering temperatures and (b) Optical micrograph (×40) of Vickers indentation of the increases from 0.43034 nm to 0.43204 nm when the sample is sintered
bulk materials sintered by SPS at 1150 °C after 20 h of milling. at 1200 °C. These calculated values (0.43034 nm and 0.43204 nm) of
lattice parameters are slightly smaller than that of TiC (0.43274 nm)
et al. [25] have produced (Ti0.8Zr0.2)C carbides prepared by MA and SPS [32]. This means that the reduction of the lattice parameter is due
at 1650 °C under 50 MPa and 5 min, characterized by high density to the smaller atomic radius (0.137 nm) of the W atom than that
(98%) and micro-hardness (HV 0.2 = 2760), which is lower than the (0.145 nm) of the Ti atom in the structure.
value obtained in this work (2978 HV). Based on these results, we Fig. 13 shows the variation of micro-hardness and relative density
can see that the increase of Vickers micro-hardness with sintering versus diffraction crystallite size of Ti0.9W0.1C carbide after SPS. The
temperature could be attributed to the hard tungsten phase. micro-hardness could be affected by both relative density and grain
However, porous materials usually have a lot of scatter. So, it is size. Pores in ceramics have no resistance to applied stress. So, materials
desirable to quantify the scatter of the measured micro-hardness values. of higher porosity have lower apparent micro-hardness than dense
The most common measure of scatter is the standard deviation. The materials or less porous materials. In addition to the effects of porosity,
Vickers micro-hardness and the standard deviation values are listed in the grain size also influences micro-hardness. In fact, smaller grain size
Table 4. The calculated standard deviation values are between 54 and increases the frequency with which dislocations encounter grain
87 HV. These values are small (about 2–3% of HV values) indicating that boundaries, thus requiring larger stress for deformation to occur [35].
the materials are quite homogeneous in microstructure. Fig. 12(b) Based on Fig. 13, one can conclude that by increasing the MA time
shows the optical micrograph of Vickers indentation carried out at from 10 h to 20 h, which leads to a decrease in the grain size, the
2.94 N on polished surface of the bulk materials sintered by SPS at micro-hardness and the relative density of the consolidated carbide
1150 °C after 20 h of milling. Based on this figure, we can see that the powders were reduced. However, as the relative density of the bulk
length of indentation is very high as compared to the pores size. In fact, material increased, the Vickers micro-hardness increased. Thus, the
micro-hardness reaches its maximum value (2978 HV) when relative
density increases to reach 98.35%. So, we assume that the micro-
Table 4 hardness is more affected by the material porosity and then by the
Calculated lattice parameters for different mechanical alloying duration (MAd) and relative density than by the grain size and the micro-hardness evolution
sintering temperature (T) under 100 MPa and 5 min, and measured Vickers micro-
doesn't follow the Hall–Petch behavior.
hardness values.
According to Song et al. [36,37], the micro-hardness of (Ti,W)C is
MAd (h) aMA (nm) TSPS (°C) aSPS (nm) HV 0.3 greater than that of TiC when the W content in (Ti,W)C is low. Indeed,
10 0.43034(2) 1200 0.43204(3) 2978 ± 54 Teber et al. [10] and Abderrazak et al. [21] have prepared TiC carbides
20 0.42947(2) 1150 0.42957(2) 2745 ± 86 by MA and SPS (5 min at 1650 °C under 100 MPa). These materials
20 0.42947(2) 1200 0.42984(1) 2846 ± 81 were characterized by 95% relative density and 2570–2700 HV micro-
20 0.42947(2) 1250 0.42904(1) 2871 ± 87
hardness values. The load used to measure TiC micro-hardness was
258 C. Slama et al. / Int. Journal of Refractory Metals and Hard Materials 43 (2014) 250–258

[3] Koc RC, Meng C, Swift GA. Sintering properties of submicron TiC powders from car-
bon coated titania precursor. J Mater Sci 2000;35:3131–41.
[4] Jin YZ, Liu Y, Wang YK, Ye JW. Study on phase evolution during reaction synthesis of
ultrafine (Ti, W, Mo, V) (CN)-Ni composite powders. Mater Chem Phys
2009;118:191–6.
[5] Kim J, Kang S. Microstructure evolution and mechanical properties of (Ti0.93W0.7)
C-xWC-20Ni cermets. Mater Sci Eng A 2011;528:3090–5.
[6] Jung J, Kang S. Sintered (Ti, W)C carbides. Scr Mater 2007;56:561–4.
[7] Jung J, Kang S. Effect of nano-size powders on the microstructure of Ti (C, N)–xWC–
Ni cermets. J Am Ceram Soc 2007;90:2178–83.
[8] Ahn S, Kang S. Dissolution phenomena in the Ti (C0.7 N 0.3)–WC–Ni system. Int J
Refract Met Hard Mater 2008;26:340–5.
[9] Sherif El-Eskandarany M. Structure and properties of nanocrystalline TiC full-density
bulk alloy consolidated from mechanical reacted powders. J Alloys Compd
2000;305:225–38.
[10] Teber A, Schoenstein F, Têtard F, Abdellaoui M, Jouini N. Effect of SPS process
sintering on the microstructure and mechanical properties of nanocrystalline TiC
for tools application. Int J Refract Met Hard Mater 2012;30:64–70.
[11] Thorne K, Ting SJ, Chu CJ, Mackenzie JD, Getman TD, Hawthorne MF. J Mater Sci
1992;27:4406.
[12] Saidi A. Reaction synthesis of Fe–(W, Ti)C composites. J Mater Process Technol
1999;89–90:141–4.
[13] Anselmi–Tamburini U, Garay JE, Munir ZA. Fast low temperature consolidation of
bulk nanometric ceramic materials. Scr Mater 2006;54:823–8.
[14] Orrù R, Licheri R, Locci AM, Cincotti A, Cao G. Consolidation/synthesis of mate-
rials by electric current activated/assisted sintering. Mater Sci Eng R
Fig. 13. Variation of Vickers micro-hardness and relative density as a function of the 2009;63:127–287.
diffraction crystallite size after sintering. [15] Abdellaoui M, Gaffet E. A mathematical and experimental dynamical phase diagram
for ball-milled Ni10Zr7. J Alloys Compd 1994;209:351–61.
[16] Abdellaoui M, Gaffet E. The physics of mechanical alloying in a planetary ball mill:
0.2 kgf. In the present study, the micro-hardness (2978 HV) and the mathematical treatment. Acta Metall Mater 1995;433:1087–98.
relative density (98.35%) of the obtained Ti0.9W0.1C are higher than [17] Rodriguez–Caravajal. Recent advances in magnetic structure determination by neu-
tron powder diffraction. J Phys B 1993;192:55–69.
those of TiC prepared by Teber et al. [10] and Abderrazak et al. [21].
[18] Rietveld HM. A profile refinement method for nuclear and magnetic structures. J
This is due to the high relative density of the material obtained after Appl Crystallogr 1969;2:65–71.
being consolidated by SPS in optimum conditions in this work. [19] Ghosh B, Pradhan SK. Microstructure characterization of nanocrystalline TiC synthe-
sized by mechanical alloying. Mater Chem Phys 2010;120:537–45.
[20] Rahaei MB, Rad RY, Kazemzadeh A, Ebadzadeh T. Mechanochemical synthesis of
4. Conclusions nano TiC powder by mechanical milling of titanium and graphite powders. Powder
Technol 2012;217:369–76.
In this work, bulk binderless Ti0.9W0.1C carbides were prepared by [21] Abderrazak H, Schoenstein F, Abdellaoui M, Jouini N. Spark plasma sintering consol-
idation of nanostructured TiC prepared by mechanical alloying. Int J Refract Met
mechanical alloying and spark plasma sintering. Based on the results Hard Mater 2011;29:170–6.
described above, the following conclusions can be drawn. [22] Chaim R, Kleiner L, Kalabukhov S. Densification of nanocrystalline TiC ceramics by
spark plasma sintering. Mater Chem Phys 2011;130:815–21.
1. The fcc Ti0.9W0.1C carbide forms since 1.5 h of MA. The formation was [23] Han Y, Fan J, Liu T, Cheng H, Tian J. The effects of ball-milling treatment on the den-
completed after 10 h of MA. sification behavior of ultra-fine tungsten powder. Int J Refract Met Hard Mater
2011;29:743–50.
2. The lattice parameter and diffraction crystallite size decrease when [24] Teber A, Schoenstein F, Têtard F, Abdellaoui M, Jouini N. The effect of Ti substitution
increasing milling time up to 10 h to reach, respectively, 0.43034 nm by Zr on the microstructure and mechanical properties of the cermet Ti1 − xZrxC
sintered by SPS. Int J Refract Met Hard Mater 2012;31:132–7.
and 8 nm. The decrease of lattice parameter of Ti0.9 W0.1C after milling
[25] Teber A, Schoenstein F, Abdellaoui M, Jouini N. Fabrication, microstructure and
time can be attributed to the formation of the Ti0.9W0.1C carbide. mechanical properties of novel bulk bindrless (Ti0.8Zr0.2)C carbides prepared
3. The increasing of milling time reduces the onset temperature of by mechanical alloying and spark plasma sintering. Ceram Int
densification. Thus, the onset temperature of densification decreases 2012;38:4929–33.
[26] Shatov AV, Firstov SA, Shatova IV. The shape of WC crystals in cemented carbides.
from 900 °C for 10 h to about 680 °C for 20 h as a result of the Mater Sci Eng A 1998;242:7–14.
smallest grain size and the most uniform microstructure. [27] Ohser-Wiedemann R, Martin U, Muller A, Schreiber G. Spark plasma sintering
4. The maximum hardness was obtained for the more dense materials, of Mo–W powders prepared by mechanical alloying. J Alloys Compd
2013;560:27–32.
meanwhile, the grain size is large. The hardness evolution doesn't [28] Ohser-Wiedemann R, Weck Ch, Martin U, Muller A, Seifert HJ. Spark plasma
follow the Hall–Petch behavior and is more affected by the material sintering of TiC particle-reinforced molybdenum composites. Int J Refract Met
porosity. Hard Mater 2012;32:1–6.
[29] Locci AM, Orrù R, Cao G, Munir ZA. Effect of ball milling on simultaneous spark plas-
5. The mono substituted Ti0.9W0.1C carbide has excellent mechanical ma synthesis and densification of TiC–TiB2 composites. Mater Sci Eng A
properties as compared to non substituted TiC carbide. Although, 2006;434:23–9.
the micro-hardness and relative density of the monosubstituted [30] Park S, Jung J, Kang S, Jeong BW, Lee CK, Ihm J. The carbon nonstoichiometry and the
lattice parameter of (Ti1 − xWx)C1 − y. J Eur Ceram Soc 2010;30:1519–26.
Ti0.9W0.1C carbide (2978 HV, 98.35%) are greater than those of TiC
[31] VitryanyukS VK, Chaplygin FI, Kostenetskaya GD. Properties of complex TiC–WC car-
carbides (2570–2700 HV, 95%) obtained by SPS sintering at almost bides in their homogeneity region. I. Preparation and phase and structural analyses
the same conditions. of TiC–WC alloys. Sov Powder Met Metall Ceram 1971;10:547–52.
[32] Powder diffraction file. Joint Committee on Powder Diffraction Standards; 1974.
[33] Biedunkiewicz A, Wrobel R. The XRD study of the nanostructured TiC/C and TiN/C
Acknowledgments composites. Rev Adv Mater Sci 2004;8:69–72.
[34] Bodziony T, Guskos N, Biedunkiewicz A, Typek J, Wrobel R, Maryniak M. Character-
ization and EPR studies of TiC and TiN ceramics at room temperature. Mater Sci
The authors would like to thank Professor M. Ben Salem and
(Poland) 2005;23:899–907.
researcher A. Kouki (Faculty of sciences Bizerte) for their valuable help. [35] Li T, Li Q, Fu JY, Yu PC, Lu L, Wu CC. Effects of AGG on fracture toughness of tungsten
carbide. Mater Sci Eng A 2007;445:587–92.
References [36] Song GM, Wang YJ, Zhou Y. Thermomechanical properties of TiC particle-reinforced
tungsten composites for high temperature applications. Int J Refract Met Hard Mater
[1] McColm IJ, Clark NJ. High performance ceramics. London: Blackie Press; 1986. 2003;21:1–12.
[2] Ellis JL, Goetzel CG. Cermets. ASM metals handbook. Ohio: ASM International; 1990 [37] Zhuzhou cemented carbide Industry Company. Manufacture of cemented carbide.
978–1007. Beijing: Chinese Metallurgical Industry Press; 1974. p. 94–7.

You might also like