You are on page 1of 10

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
Materials Science and Engineering A 527 (2010) 68706878
Contents lists available at ScienceDirect
Materials Science and Engineering A
j our nal homepage: www. el sevi er . com/ l ocat e/ msea
Microwave and conventional sintering of 90W7Ni3Cu alloys with premixed
and prealloyed binder phase
Avijit Mondal
a
, Anish Upadhyaya
a,
, Dinesh Agrawal
b
a
Department of Materials Science and Engineering, Indian Institute of Technology Kanpur, Kanpur 208016, India
b
The Pennsylvania State University, University Park, PA 16802, USA
a r t i c l e i n f o
Article history:
Received 9 April 2010
Received in revised form 8 June 2010
Accepted 22 July 2010
Keywords:
Microwave sintering
Conventional sintering
Tungsten heavy alloys
a b s t r a c t
The present study investigates the possibility of consolidating premixed 90W7Ni3Cu alloy desig-
nated as 90WPM (NiCu) through microwave sintering. An attempt has been made to compare the
results between microwave and conventionally sintered samples. This study also compares the sintering
behavior of 90W7Ni3Cu with prealloyed 90WPA (NiCu) in both conventional as well as microwave
furnace at various temperatures. The comparative analysis is based on the sintered density, densica-
tion parameter, hardness and microstructures of the samples. The present investigation also includes the
variation of matrix composition as a function of temperature by EPMA analysis. The results show that
microwave sintering requires about 75% less processing time than required by conventional method and
still provides better physical and mechanical properties.
2010 Elsevier B.V. All rights reserved.
1. Introduction
Tungsten heavy alloy (WHA) is a group of two-phase
composites, based on WNiCu and WNiFe alloys.
Tungstennickelcopper alloys are widely used for ordinance
application, electrical contacts of switches, radiation shielding,
mass balances, etc. In general the tungsten heavy alloys have been
processed through powder metallurgy route since 1930s [1]. These
heavy alloys contain mainly pure tungsten as principal phase in
association with a binder phase containing transition metals (Ni,
Fe, Cu, Co) [2]. In WNiCu alloys, normally the nickel-to-copper
ratio ranges from3:2 to 4:1. Price et al. [3] were the rst to propose
NiCu as the binder for tungsten heavy alloys. Over the last several
years, these alloys have been extensively investigated for den-
sication mechanism, microstructural evolution and properties
[47].
The effect of tungsten and copper powder size variation on
the sintered properties of WNiCu heavy alloys was carried out
by Srikanth and Upadhyaya [8,9]. The effect of composition and
sintering temperature on the densication and microstructure of
WNiCu heavy alloys was studied by Ramakrishnan and Upad-
hyaya [10]. Kuzmic [11] proposed that rapid cooling from the
sintering temperature prevents the formation of brittle phase in

Corresponding author at: Materials Science and Engineering, Indian Institute of


Technology Kanpur, P.O. IIT Kanpur, Kanpur 208016, UP, India.
Tel.: +91 512 2597672; fax: +91 512 2597505.
E-mail address: anishu@iitk.ac.in (A. Upadhyaya).
order to obtain good mechanical properties. Ariel et al. [12] corre-
lated the mechanical properties of WNiCu system with sintered
microstructure. Their study showed that the mechanical properties
are a function of mean free path between tungsten grains, volume
fractionof tungstengrains andthecontiguityof tungstenspheroids.
The solubility of tungsten in the liquid binder plays a domi-
nant role in determining the mechanical properties of the sintered
WNiCu alloys. The solubility of tungsten in copper is negli-
gible. Even at temperatures as high as 1350

C only 0.04at.% of
tungsten goes in solution with copper [13]. In contrast, tungsten
exhibits appreciable solubility (up to 40wt.%) in nickel. It is there-
fore possible to tailor the tungsten solubility, wetting and the
dihedral angle and thereby, the accompanying sintering response
and properties of the system by varying the Ni:Cu ratio [10,14].
Nowadays, WNiCu alloys are also being consolidated by employ-
ing prealloyed powders [15]. Use of prealloyed powder improves
homogeneity. The homogenization process accelerates sintering
and promotes densication. The alloy formation during sintering
decreases the material viscosity and, hence, stimulates material
ow under the action of capillary forces [16].
Despite widespread application, difculties still exist in the
manufacture of liquid phase sintered tungsten heavy alloys. In
order toavoidthermal shock, processingof tungstenheavyalloys in
conventional furnace involves heating at a slower rate (<10

C/min)
and with isothermal holds at intermittent temperatures. This not
only increases the process time, but also results in signicant
microstructural coarsening during sintering, leading to the degra-
dationof mechanical properties. This problemis further aggravated
when the initial powder size is extremely ne. Hence, it is envis-
0921-5093/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2010.07.074
Author's personal copy
A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878 6871
Table 1
Powder characteristics of as-received elemental W, Ni, Cu and prealloyed NiCu powder.
Property W Ni Cu 69Ni31Cu
Supplier Widia Inco A Cu Powder International, LLC Ametek
Processing technique Chemical reduction Carbonyl process Gas atomization Gas atomization
Powder shape Irregular Spiky Spherical Spherical
Powder size (m)
D
10
2.0 3.8 19.4 8.2
D
50
4.2 11.0 47.0 17.3
D
90
6.2 31.8 96.3 40.4
aged that a fast heating rate would mitigate this problem. One of
the techniques to achieve fast and relatively uniform sintering is
through microwaves [17]. The WNiCu alloys are usually pre-
pared by mixing the constituent powders. One of the challenges
in the processing of these alloys is to ensure homogenous melt
distribution [15]. Nickel and copper are known to interdiffuse at
temperatures higher than 900

C. However, in case of rapid sinter-


ing in microwave this may not be readily feasible.
Microwave heating is much more uniformat a rapid rate result-
ing in reduction of processing time and energy consumption. Also
rapid heating leads to ner microstructure enhancing the mechan-
ical properties. Clark and Sutton [18] cited many other benets of
the process such as precise and controlled heating, environmen-
tally friendly etc. Microwave heating is a very sensitive function
of the material being processed and depends on several factors,
such as sample size, as-pressed density, its mass and geometry
[19,20]. Though there have been attempts to explain microwave
heating of metal powders, still there is not yet any consensus on a
comprehensive theory to explain the mechanism [21].
Applicability of microwave sintering to metals was ignored due
to the fact that they reect microwaves. Roy et al. [22] reported
that particulate metals can be heated rapidly in microwaves. This
has led to the use of microwaves to consolidate a range of partic-
ulate metals and alloys [2325]. While researchers have reported
microstructural renement duetorapidheatinginmicrowaves, the
effect of microwave sintering on the microstructural homogene-
ity and mechanical properties seem to be system specic [2325].
Researchers have also attempted to consolidate refractory materi-
als such as pure W, WCu, WNiFe and WNiCu [2633] using
microwave energy and reported signicant reduction in process
time, elimination of brittle intermetallic formation and superior
mechanical properties.
This paper reports the sintering behavior of premixed and pre-
alloyed 90W7Ni3Cu powders in both, microwave as well as
in a conventional (radiatively heated) furnace. The comparative
analysis is based on the sintered density, densication parameter,
hardness and microstructures of the samples.
2. Experimental procedure
The as-received powders were characterized for their size, size
distribution and morphology. Table 1 summarizes the charac-
teristics of the as-received elemental W, Ni, Cu and prealloyed
NiCu powder used for this study. For investigating the densi-
cation response and compaction studies of various compositions,
cylindrical pellets (diameter: 16mm and height 8mm) were
pressedat 200MPausingauniaxial semi-automatic hydraulic press
(model: CTM 50, supplier: FIE, India) of 50T capacity. To facilitate
compaction and subsequent removal of compacted samples, zinc-
stearate powder was applied as die-wall lubricant. The as-pressed
green compact density varied from 56% to 58% of the theoretical
density of the alloy, where the theoretical density was calculated
using the inverse rule of mixtures.
Tostudythedensicationbehavior; thegreen(as-pressed) com-
pacts were sintered using conventional and microwave furnace.
The experimental details have been mentioned elsewhere [32].
The sintered density was obtained by both dimensional mea-
surements as well as Archimedes density measurement technique.
To compare the densicationresponse of various compositions, the
sintered densities were normalized with respect to the theoretical
density. To take into account the inuence of the initial as-pressed
density, the compact sinterability was also expressed in terms of
densication parameter which is calculated as follows:
densication parameter=
sintered density green density
theoretical density green density
(1)
An Anter 1161V vertical dilatometer was used for measuring
axial shrinkage and shrinkage rate during sintering with constant
heating rate of premixed and prealloyed 90W7Ni3Cu compacts.
It measures dimensional changes over the entire sintering cycle
with precision of 1m. The dilatometery studies were performed
at constant heating and cooling rate of 10

C/min. The sintered


samples were wet polished in a manual polisher (model: Lunn
Major, supplier: Struers, Denmark) using a series of 6m, 3mand
1m diamond paste, followed by cloth polishing using a 0.04m
colloidal SiO
2
suspension. The scanning electronmicrographs of as-
polished samples were obtained by a scanning electronmicroscope
(model: FEI quanta, Netherlands) in the secondary electron (SE)
mode. Aquantitative analysis was also carriedout onselectedspec-
imens using EPMA equipment (model: JXA-8600 SX Super-Probe,
supplier: JEOL, Japan). Phase determination and phase evolution, if
any were studied for all the samples using an X-ray Diffractometer
(model: Rich. Seifert & Co., GmbH & Co., KG, Germany).
Bulk hardness measurements were performed on polished sur-
faces of sintered cylindrical compacts at a load of 5kg using Vickers
hardness tester (model: V100-C1, supplier: Leco, Japan). The load
was applied for 30s. Micro hardness tester (model: 8299, supplier:
Leitz, Germany) was used to evaluate the hardness of the matrix
phase. The load applied on matrix phase during the micro hard-
ness measurement was 15g. The bulk hardness of the compact
and the micro hardness of the matrix phase for each sample were
an average of 5 readings taken at different locations on respec-
tive phases. Transverse rupture strength (TRS) measurements of
premixed and prealloyed samples were performed following the
procedure described in MPIF Standard 41 [34].
3. Results and discussion
3.1. Heating response and densication of WNiCu alloy
Fig. 1 compares the thermal proles of 90W7Ni3Cu compacts
(prepared using premixed and prealloyed NiCu binder) in a con-
ventional and microwave furnace. It is evident fromthe gure that
WNiCu alloys couple with microwaves and undergo rapid heat-
ing. In case of conventional furnace, in order to ensure uniform
heating, the compacts heating rate was restricted to 5

C/min and
isothermal holds were provided at intermittent temperatures. In
Author's personal copy
6872 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878
Fig. 1. Thermal proles of 90W7Ni3Cu alloys heated up to 1450

C in conven-
tional and microwave furnace. The heating prole of WNiFe alloys prepared using
premixed and prealloyed NiCu binder is shown separately in case of microwave
processing.
contrast, the overall heating rate achieved in the microwave fur-
nace was 22

C/min. Takingintoconsiderationthe slowheatingrate


and isothermal holds associated with conventional sintering, the
compacts were consolidated through microwaves with signicant
(75%) reductionintheprocessingtime. Figs. 2and3showthepho-
tograph of the 90W7Ni3Cu alloys sintered in the temperature
range of 12001450

C in conventional and microwave furnace,


respectively. Note that irrespective of the heating mode and the
binder chemistry, none of the 90W7Ni3Cu compacts distorted
or developed any cracks during sintering even up to temperatures
as high as 1450

C. It is interesting to note that despite a fast heat-


ing rate, no macro cracking was observed in any of the microwave
sintered samples. This is attributed to the volumetric heating of
the compacts, a unique feature of microwave heating. From the
microwaveheatingprole(Fig. 1), it is interestingtonotethat alloys
with premixed NiCu binder heat at slightly faster rate than those
containing prealloyedmatrix. Elsewhere, Sethi et al. [23] andUpad-
hyaya and Sethi [35] too have reported similar heating response of
the premixed and prealloyed Cu12Sn bronze in microwave fur-
Fig. 2. Photograph of 90W7Ni3Cu compacts conventionally sintered at various
temperatures ranging from 1200

C to 1450

C.
Fig. 3. Photograph of the 90W7Ni3Cu compacts prepared using (a) prealloyed
and (b) premixed NiCu binder and sintered in a microwave furnace at various
temperatures.
nace. It can therefore be inferred that the chemical state of the
alloying constituents inuences the microwave coupling and the
resultant heating. It is well recognized that microwave heating
is material dependent. Peelamedu et al. [36] and Roy et al. [37]
demonstrated in multi-phase systems, the phases heat up at dif-
ferent rates and lead to local an isothermal heating effect. Through
model experiments in systems, such as Y
2
O
3
Fe
3
O
4
, BaCO
3
Fe
3
O
4
,
and NiOAl
2
O
3
, Peelamedu et al. [36] demonstrated that the hotter
Fig. 4. Effect of binder state (premixed vs. prealloyed), heating mode and sintering temperature on the (a) sintered density and (b) densication parameter of 90W7Ni3Cu
alloys. For the sake of comparison, the density and densication parameter data points (denoted by symbol ) for the same composition (in premixed state) sintered at
1300

C and 1400

C are also superimposed.


Author's personal copy
A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878 6873
Table 2
Composition as a function of temperature and heating mode of the matrix phase of liquid phase sintered 90W7Ni3Cu alloys prepared by (a) premixed and (b) prealloyed
powders.
(a) Premixed powders
Temperature (

C) Heating mode Elemental content (wt.%)


W Ni Cu
1200 Conventional 19.13 1.93 53.15 2.71 25.69 0.92
Microwave 18.43 0.37 55.03 0.85 24.18 0.61
1250 Conventional
Microwave 20.41 0.72 56.16 0.89 20.77 0.52
1300 Conventional 18.66 0.42 56.04 0.68 23.98 0.88
Microwave 24.29 4.65 54.01 1.84 18.37 5.72
1450 Conventional 22.06 4.66 57.03 1.02 20.27 5.36
Microwave 23.57 3.37 59.36 0.35 14.41 3.09
(b) Prealloyed powders
Temperature (

C) Heating mode Elemental content (wt.%)


W Ni Cu
1200
Conventional 29.25 0.73 53.62 1.00 15.59 0.13
Microwave 27.26 0.96 54.42 0.65 15.20 0.40
1250
Conventional
Microwave 25.12 0.57 55.68 0.78 16.32 0.55
1300
Conventional 26.37 0.38 56.36 0.39 15.36 0.22
Microwave 28.73 18.0 50.34 6.99 17.08 10.63
1450
Conventional 23.25 7.56 56.96 2.3 18.80 5.60
Microwave 19.34 3.09 57.97 2.71 20.62 5.80
species diffuse rapidly intothe relatively colder ones andcorrelated
it with the very rapid diffusion rates observed in microwave pro-
cessed compacts. It is likely that a similar mechanismis also active
in the W heavy alloys system as well. The presence of such effect
will result in densication enhancement in the microwave sintered
components.
To test the above hypothesis, the densication responses of the
compacts consolidated at various temperatures was evaluated for
both the heating modes and is summarized in Fig. 4a and b. From
Fig. 4a, it is evident that the density of the compacts increases
with increasing sintering temperature. Since the green density
of all the compacts were the same, the densication parameter
results (Fig. 4b) closely follow the same trend as sintered den-
sity. Note that with the exception of 1200

C, for all other sintering


temperatures, microwave sintering results in better densication.
Since factors that contribute to enhanced diffusivity are known to
enhance densication during sintering, it can be inferred that the
atomic mass transport is higher incase of microwave sintering. This
will also result into less grain coarsening during microwave sinter-
ing. Consequently, the ner particle size is retained at the sintering
temperature which results in higher grain boundary diffusivity.
From the sintered density and the densication parameter
plots shown in Fig. 4, it is evident that for both heating modes
the densication increases with increasing temperature. For the
sake of comparison the sintered density and densication param-
eter results reported by Ramakrishnan and Upadhyaya [10] on
the similar alloy composition sintered at 1300

C and 1400

C has
been superimposed on Fig. 4a and b. Note that the densication
results from this study on 90WPM (NiCu) are similar to those
reported by Ramakrishnan and Upadhyaya [10]. Based on the sin-
tering studies on a range of WNiCu alloys, Makipiritti [38] and
Kothari [6] have conrmed that densication in this system is
diffusion-controlled. Kothari [6] and Crowson [39] have shown
that increasing sintering temperature results in enhanced diffu-
sivity and lower activation energy in WNiCu alloys. The same
will hold good for the present investigation as well and thereby
activate densication. Besides enhanced diffusivity at high tem-
perature, another factor that contributes to densication is the
solid-solubility of tungsten in the NiCu matrix [3840]. In one of
the veryearlyworks, Prill et al. [41] basedonthe sinteringstudies of
99W1(NiCu) alloys showed that the dissolution of the tungsten
grains in the matrix and their subsequent reprecipitation from the
supersaturated matrix on to the larger tungsten is a major con-
tributor to densication. To validate this, the effect of sintering
temperature, heating mode and the powder type on the matrix
composition was evaluated using EPMA. The corresponding results
for 90WPM (NiCu) and 90WPA (NiCu) compacts are summa-
rized in Table 2a and b, respectively. From Table 2a, it is obvious
that higher sintering temperature leads to higher tungsten solubil-
ity in the matrix for the samples prepared from premixed powder.
Thematrixcompositionanalysis is similar tothosereportedbySail-
landet al. [42] andXiong et al. [43] for the sinteredWNiCualloys.
On the other hand for the prealloyed compositions (Table 2b) the
tungsten solubility does not follow any specic trend as a function
of temperature. The complete explanation for this observation is
not readily available. It is hypothesized that this trend is possibly
due to the difference in diffusivity of tungsten in prealloyed NiCu
matrix.
3.2. Dilatometric evaluation of 90W7Ni3Cu alloy
For better understanding the phenomenology of the densi-
cation of the WNiCu compacts the dimensional changes were
measured in situ during sintering using dilatometry. However,
owing to the instruments limitations, dilatometric evaluation
could be performed through conventional heating mode only.
Fig. 5a and b plots the dilatometric data of shrinkage and shrinkage
rate vs. temperature for the as-pressed 90W7Ni3Cu compacts
prepared using premixed and prealloyed binder. From the gures,
it is evident that both compacts undergo gradual densication
during heating till 1280

C. Subsequent heating beyond 1280

C
results insuddenincrease inthe axial shrinkage andshrinkage rate.
Author's personal copy
6874 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878
Fig. 5. Dilatometer plots showing the effect of NiCubinder state (premixedvs. pre-
alloyed) on the (a) shrinkage and (b) shrinkage rate of 90W7Ni3Cu alloys heated
up to 1400

C at a constant heating rate of 10

C/min in reducing atmosphere.


This spurt in the densication is attributed to the onset of binder
melting which results in densication through capillary-induced
rearrangement and solution-reprecipitation. The rapid shrink-
age resulting from liquid phase sintering sustains until 1400

C
wherein nearly full densication (98%) was achieved. Elsewhere,
Wu et al. [44] and Martin et al. [15] too have reported simi-
lar dilatometric curves for 80W14Ni6Cu alloys and prealloyed
90W6.9Ni3.1Fe, respectively. For the range of compositions
used, both researchers [15,44] have demonstrated that the onset
of melt formation in the WNiCu system with Ni and Cu in
the ratio 7:3 occurs at 1282

C. From the binary phase dia-


gram [45] and the experiments [44] the solidus and the liquidus
temperatures of the 70Ni30Cu matrix was determined to be
1345

C and 1380

C, respectively. However, in case of ternary


WNiCu system, due to substantial solubility of the tungsten
in the matrix the solidus and liquidus temperatures are lowered
[42].
From the dilatometric results (Fig. 5a and b), it can be dis-
cerned that the 90WPA (NiCu) alloy exhibits slightly higher
densicationparticularlyat lower temperatures as comparedtothe
90WPM (NiCu) compact. This trend is similar to that reported
by Huang et al. [46] who investigated the inuence of preal-
loyed NiFeMo binder on the sintering behavior of tungsten
heavy alloys. They [46] noted that prealloying results in a more
Fig. 6. XRD plots showing the effect of heating mode and sintering temperature on
the phase evolution of 90W7Ni3Cu alloys prepared using (a) premixed and (b)
prealloyed NiCu binder.
homogenous microstructure and lowers the interfacial energy of
the tungsten (solid)matrix (liquid) phase providing a higher den-
sity compact. The analyses of matrix composition of the WNiCu
compacts in the present study (Table 2a and b) particularly, the
ones sintered at 1200

C are in accordance with this hypothe-


sis. At 1200

C, the dissolved tungsten amount in 90WPM(NiCu)


is lower than its prealloyed counterpart. Consequently, the lat-
ter exhibits higher densication (Fig. 4b). For compacts heated or
sintered at higher temperatures, the densication in both the com-
pacts is not too different and is attributed to homogenization of
the premixed matrix through interdiffusion. It is well known that
Cu and Ni have mutual intersolubility [45]. Using X-ray analysis,
Heckel et al. [47] have experimentally shown that signicant inter-
diffusion and homogenization through in situ alloying occurs in the
Ni52at.% Cu compact after sintering at 950

C for 1h. Since the


Cu:Ni ratio in the present system (90W7Ni3Cu) is much lower,
hence, it is quite reasonable to assume that the premixed NiCu
powders will get completely homogenized during conventional
sintering.
3.3. Microstructural evolution and properties
Fig. 6a and b compares the effect of heating mode and
sintering temperature on the phase evolution in 90WPM
Author's personal copy
A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878 6875
Fig. 7. SEM micrographs of 90W7Ni3Cu alloys prepared using prealloyed NiCu binder and consolidated in a microwave furnace at (a) 1200

C, (b) 1250

C, (c) 1300

C
and (d) 1450

C, respectively.
(NiCu) and 90WPA (NiCu) alloys, respectively. For all
XRD experiments, the sample size and other test parame-
ters were kept constant. It is therefore rather surprising to
note that in case of microwave sintered compacts some of
the peaks do not even register on the diffractogram. Sim-
ilar observations have also been reported by Padmavathi
[48] in case of microwave sintering of various aluminium
alloys.
The reduction in the intensity indicates that the microwave
sintered compacts have less perfect crystals and contain more
defects as compared to their conventionally sintered counter-
parts. Recently, many researchers, [37,4951] have reported that
Fig. 8. SEM micrographs of (a) premixed and (b) prealloyed 90W7Ni3Cu alloys sintered at 1300

C in conventional (left) and microwave (right) furnace.


Author's personal copy
6876 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878
Fig. 9. SEM micrographs of (a) premixed and (b) prealloyed 90W7Ni3Cu alloys sintered at 1450

C in conventional (left) and microwave (right) furnace.


exposure to microwave heating leads to decrystallization and
sometimes even amorphization in some of the ceramic and semi-
conducting systems. The defects within a crystalline structure are
known to activate sintering [52]. Hence, it may be possible that
in the present system too the defects generated on exposure to
microwaves contribute to densication enhancement.
The phenomenology of microstructural evolution in conven-
tionally sinteredWNiCualloys has beenextensively investigated
by several researchers [3,10,15,40]. In the present study, the atten-
tion was focused on the microstructural response of microwave
sintered WNiCu system and its comparison with that obtained
throughconventional sintering. Fig. 7adshows the effect of sinter-
ing temperature on the representative microstructure of 90WPA
(NiCu) alloyconsolidatedina2.45GHzmultimodemicrowavefur-
nace. It is quite evident that higher temperature not only results in
pore elimination but is also accompanied by redistribution of the
matrix and coarsening of tungsten grains. It is worth noting that
the microstructure of the compacts sintered in solid-state (1200

C
and 1250

C) is drastically different from the liquid phase sintered


counterparts (Fig. 7c andd). Incase of compacts sinteredat 1200

C,
the binder phase is inhomogenously distributed (Fig. 7a) and the
tungsten grains are faceted. In contrast, at 1250

C (Fig. 7b), while


the W-grains still remain faceted, the binder is more uniformly dis-
tributed. Thewell roundedandcoarser grains associatedwithliquid
phase sintering (Fig. 7c and d) are associated with higher solid-
solubility (Table 2) and relatively faster solution-reprecipitation
kinetics.
Fig. 8a and b compares the effect of heating mode on the
microstructures of the 90WPM (NiCu) and 90WPA (NiCu)
alloys, respectively, sintered at 1300

C. The corresponding
microstructures of compacts liquid phase sintered at 1450

C are
shown in Fig. 9a and b. From Figs. 8 and 9, it can be discerned that
irrespective of the sintering temperature and the heating mode,
the prealloyed sintered compacts have slightly higher coarsening
than their premixed counterparts. This can be attributed to a more
homogenous microstructure and better spreading of the melt due
to the relatively lower tungsten (solid)matrix (liquid) interfacial
energyincase of alloys preparedusingprealloyedbinder [46]. More
recently, Martin et al. [15] showed that the NiCu additive phase
in 90W6.9Ni3.1Cu alloy with prealloyed matrix is more effec-
tively distributed within the tungsten grains during sintering and
contributes not only to densication but also leads to enhanced
tungsten grain coarsening as compared to premixed alloys.
One of the intriguing observations in the present study is the
dramatic difference between the microstructure of WNiCu com-
pacts sintered in conventional and microwave furnace (Fig. 8a and
b). In case of conventionally sintered compacts, the microstructure
at 1300

C is quite similar to that obtained through solid-state sin-


tering. Some of the earlier researchers [10,53,54] too have reported
similar ndings at this temperature for premixed 90W7Ni3Cu
compacts. The larger tungsten grain size in microwave sintered
WNiCu compacts at 1300

C is contrary to the observations in


various systems that afast heatingratewill result inlower grainsize
[30,55,56]. However, similar contrary results have recently been
reported by Zaspalis et al. [57] in microwave sintered MnZn fer-
rites and by Kim et al. [58] in spark plasma sintered alumina. The
exact explanation of such observations is still lacking and further
detailed experiments need to be carried out in future to elucidate
this better. In contrast to 1300

C sintered compacts, both con-


ventional as well as microwave sintered 90W7Ni3Cu alloys (at
1450

C) exhibit well-dened liquid phase sintered microstructure


with rounded tungsten grains interspersed in the NiCu matrix
(Fig. 9a and b). Table 3 summarizes the effect of heating mode
on the average tungsten grain size in premixed and prealloyed
90W7Ni3Cu alloys liquid phase sintered at 1450

C. From Fig. 9
and Table 3, it can be inferred that irrespective of the state of ini-
Author's personal copy
A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878 6877
Table 3
Effect of heating mode on the average tungsten grain size in 90W7Ni3Cu alloys
prepared using premixed and prealloyed NiCu binder and sintered at 1450

C.
Composition Heating mode W grain size (m)
90WPM
(NiCu)
Conventional 23 10
Microwave 21 7
90WPA
(NiCu)
Conventional 26 8
Microwave 20 7
Table 4
Effect of heating mode andsintering temperature onthe bulk hardness (inkgf/mm
2
)
of 90W7Ni3Cu alloys prepared using premixed and prealloyed binder.
Heating mode 90WPM (7Ni3Cu)
a
90WPA (7Ni3Cu)
1300

C 1450

C 1300

C 1450

C
Conventional 276 21 301 16 290 19 330 13
Microwave 309 17 328 13 330 18 350 12
a
As a note of comparison in one of the previous reports [10] the bulk hardness of
the same compositionhas beenlistedas 175VHNand220VHNfor compacts sintered
at 1300

C and 1400

C, respectively.
tial binder chemistry (premixed vs. prealloyed), the microwave
sintered compacts have relatively more rened microstructure
as compared to those consolidated in a conventional (radiatively
heated) furnace. Elsewhere, Upadhyaya et al. [30] has reported
similar ndings in liquid phase sintered WNiFe system.
The differences in the sintered microstructure of alloys consol-
idated through conventional and microwave furnace is bound to
inuence the mechanical properties. Table 4 summarizes the effect
of heating mode and powder chemistry (prealloyed vs. premixed)
on the bulk hardness of 90W7Ni3Cu alloys sintered at 1300

C
and 1450

C, respectively. To provide a basis of comparison, the


bulk hardness of premixed 90W7Ni3Cu alloy sintered at 1300

C
and 1400

C reported by Ramakrishnan and Upadhyaya [10] is also


provided in Table 4. It is remarkable to note that the hardness
achieved in the present study is far superior to those reported in
earlier work [10,59]. From Table 4, it is worth noting that for all
conditions, microwave sintering results in up to 13% improvement
in the bulk hardness. This can be attributed to the increased densi-
cation and a more rened microstructure in case of microwave
sintered alloys. Previous report from Upadhyaya et al. [30] has
shown that the hardness of tungsten grains per se does depend
on the heating mode. However, on account of the varying tungsten
solid-solubility, the hardness of the matrix phase is expected to
be dependent upon the heating mode. To test this, Table 5 summa-
rizes the corresponding matrix hardness measurements performed
oncompacts consolidatedinboththe heating modes at 1300

Cand
1450

C. It is interesting to observe that in case of compacts con-


solidated at 1300

C, the matrix hardness is higher for microwave


processed compacts (Table 5). This can be correlated to the higher
tungstensolubility observedinmicrowave sinteredalloys (Table 2a
and b). In contrast, the trend is reversed for compacts sintered at
1450

C. In case of prealloyed compacts, this can be correlated to


the higher W-solubility inconventionally sinteredalloys. However,
further tests need to be performed to gain detailed insight for this.
Table 5
Effect of heating mode on the micro hardness (in kgf/mm
2
) of the matrix phase in
sintered 90W7Ni3Cu alloys.
Heating mode 90WPM (7Ni3Cu) 90WPA (7Ni3Cu)
1300

C 1450

C 1300

C 1450

C
Conventional 260 20 333 13 250 25 320 16
Microwave 290 23 312 15 280 21 290 12
Fig. 10. Effect of heating mode and sintering temperature on the transverse rupture
strength of 90W7Ni3Cu alloys prepared using premixed and prealloyed binders.
Fig. 10 shows the effect of heating mode and powder chem-
istry on the transverse rupture strength of the 90W7Ni3Cu alloy
sinteredat 1300

Cand1450

C. Oneof theveryinterestingobserva-
tions is that microwave sintered compacts have higher transverse
rupture strengthas comparedto those sinteredconventionally. The
enhancement of strength is as high as up to threefolds (211%) in
case of alloys processed at 1300

C. This can be correlated to a less


contiguous microstructure consisting of rounded tungsten grains
that result in alloys microwave sintered at 1300

C (Fig. 8). Else-


where, Chermant and Osterstock [60] have demonstrated similar
inuence of microstructure on the fracture toughness of WCCo
system. The transverse rupture strength of the conventionally sin-
tered 90WPM (NiCu) system is similar to that reported earlier
by Kalina et al. [61]. Another observation that can be inferred from
Fig. 10 is that the transverse rupture strength of alloys prepared
with prealloyed NiCu is marginally higher than those prepared
using premixed binder. This could be due to better homogeneity of
the alloying additives in case of the prealloyed system.
4. Conclusions
90W7Ni3Cu alloys, compacts were prepared using both
premixed and prealloyed NiCu binder and were sintered at
temperatures ranging between 1200

C and 1450

C. As com-
pared to conventional sintering, both premixed and prealloyed
90W7Ni3Cu alloys were microwave sintered in signicantly
(75%) less time. In spite of higher heating rate in microwave sin-
tering, no micro- or macro-cracking was observed in all microwave
sintered samples. Dilatometric analysis of the 90W7Ni3Cu alloy
at slower heating rates (<10

C/min) reveals that unlike the WCu


system the WNiCu system exhibits signicant solubility of
tungsteninthe matrix. Consequently, a major portionof the overall
densication in this alloy occurs prior to melt formation. Due to the
inter-diffusivity between nickel and copper, by the time the sinter-
ing temperature was attained, both Cu and Ni forma homogeneous
solid solution. Hence, the sintering response of the 90W7Ni3Cu
alloy was per se independent of the initial matrix type (premixed
vs. prealloyed). An interesting nding in the present study was the
observation of the heating mode effect on the microstructure of the
compacts sintered at 1300

C. While the conventionally processed


alloy had a typical solid-state sintered structure, the microwave
processed samples had a well-developed microstructure, typically
associated with liquid phase sintering. Consequently, as compared
Author's personal copy
6878 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 68706878
to conventional sintering, microwave sintering resulted in about
9% improvement in hardness (to 350VHN) and about threefold
(211%) enhancement in exural strength to around 1625MPa.
Acknowledgement
This collaborative research was done as a part of the Center for
Development of MetalCeramic Composites through Microwave
Processing which was funded by the Indo-US Science and Tech-
nology Forum (IUSSTF), New Delhi.
References
[1] S.G. Caldwell, in: P.W. Lee, R. Lacocca (Eds.), Powder Metal Technologies and
Applications, vol. 7, Materials Park, OH, USA, 1998, pp. 914921.
[2] A. Upadhyaya, Mater. Chem. Phys. 67 (2001) 101110.
[3] G.H.S. Price, C.J. Smithells, S.V. Williams, J. Inst. Met. 62 (1938) 239264.
[4] A.L. Prill, Trans. AIME 230 (1964) 769772.
[5] J.H. Brophy, H.W. Hayden, J. Wulff, Trans. TMS-AIME 224 (1962) 797803.
[6] N.C. Kothari, J. Less Common Met. 13 (1967) 457468.
[7] R. Makarov, O.K. Teodorovich, I.N. Fruntsevich, Sov. Powder Metall. Met. Ceram.
4 (1965) 554559.
[8] V. Srikanth, G.S. Upadhyaya, Int. J. Ref. Hard Met. 2 (1983) 4954.
[9] V. Srikanth, G.S. Upadhyaya, Powder Technol. 39 (1984) 6165.
[10] K.N. Ramakrishnan, G.S. Upadhyaya, J. Mater. Sci. Lett. 9 (1990) 456459.
[11] J.F. Kuzmic, in: H.H. Hausner (Ed.), ModernDevelopment inPowder Metallurgy,
vol. 3, Plenum Press, New York, 1966, pp. 166171.
[12] E. Ariel, J. Batra, D. Brandon, Powder Metall. Int. 5 (1973) 125128.
[13] V.N. Eremenko, R.V. Minakova, M.M. Churakov, Sov. Powder Metall. Met. Ceram.
15 (1976) 283286.
[14] Y.V. Naidich, I.A. Lavrinenko, V.A. Evdokimov, Sov. Powder Metall. Met. Ceram.
16 (1977) 276280.
[15] J.M. Martin, J.L. Johnson, R.M. German, F. Castro, in: E. Daver, C.J. Trombino
(Eds.), Advances in Powder Metallurgy and Particulate Materials, vol. 1, Metal
Powder Industries Federation, Princeton, NJ, 2006, pp. 5.435.57.
[16] A.P. Savitskii, Sci. Sinter. 37 (2005) 317.
[17] K.J. Rao, P.D. Ramesh, Mater. Bull. Sci. 18 (1995) 447465.
[18] D.E. Clark, W.H. Sutton, Ann. Rev. Mater. Sci. 26 (1996) 299331.
[19] R.M. Hutcheon, M.S. De Jong, F.P. Adams, Microw.: Theor. Appl. Mater. Proc. III,
Ceram. Trans., Am. Ceram. Soc. 59 (1995) 215.
[20] A. Mondal, D. Agrawal, A. Upadhyaya, J. Microw. Power. Electromagn. Energy
(JMPEE) 43 (2009) 510.
[21] M.W. Porada, Microw.: Theor. Appl. Mater. Proc. III, Ceram. Trans., Am. Ceram.
Soc. 80 (1997) 153.
[22] R. Roy, D.K. Agrawal, J.P. Cheng, S. Gedevanishvili, Nature 399 (1999) 668670.
[23] G. Sethi, A. Upadhyaya, D. Agrawal, Sci. Sinter. 35 (2003) 4965.
[24] R.M. Anklekar, D.K. Agrawal, R. Roy, Powder Metall. 44 (2001) 355362.
[25] K. Saitou, Scripta Mater. 54 (2006) 875879.
[26] M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding, D. Kapoor, D. Agar-
wal, J. Chang, Int. J. Powder Metall. 42 (2006) 4550.
[27] G. Prabhu, A. Charkraborty, B. Sarma, Int. J. Ref. Met. Hard Mater. 27 (2009)
545548.
[28] S.D. Luo, J.H. Yi, Y.L. Guo, Y.D. Peng, L.Y. Li, J.M. Ran, J. Alloys Compd. 473 (2009)
L5L9.
[29] A. Mondal, A. Upadhyaya, D. Agrawal, J. Microw. Power Electromagn. Energy
(JMPEE) 43 (2009) 1116.
[30] A. Upadhyaya, S.K. Tiwari, P. Mishra, Scripta Mater. 56 (2007) 58.
[31] A. Mondal, D. Agrawal, A. Upadhyaya, J. Microw. Power Electromagn. Energy
(JMPEE) 44 (2010).
[32] A. Mondal, D. Agrawal, A. Upadhyaya, Mater. Sci. Technol. (2008) 25022515.
[33] A. Mondal, Modeling of microwave heating of particulate metals and its appli-
cation in sintering of tungsten-based alloys, Ph.D. Thesis, The Indian Institute
of Technology, Kanpur, 2009.
[34] MPIF Standard 41, Determination of transverse rupture strength of powder
metallurgy materials, standard test methods for metal powders and powder
metallurgy products, MPIF, Princeton, NJ, USA, 2002, pp. 5557.
[35] A. Upadhyaya, G. Sethi, Scripta Mater. 56 (2007) 469472.
[36] R.D. Peelamedu, R. Roy, D. Agrawal, Mater. Res. Bull. 36 (2001) 27232739.
[37] R. Roy, Y. Fang, J. Cheng, D.K. Agrawal, J. Am. Ceram. Soc. 88 (2005) 16401642.
[38] S. Makipiritti, Powder Metall. 4 (1961) 97111.
[39] A. Crowson, Modern Developments in Powder Metallurgy, vol. 15, Metal Pow-
der Industries Federation, Princeton, NJ, 1985, pp. 507520.
[40] G. Sethi, S.J. Park, J.L. Johnson, R.M. German, Int. J. Ref. Met. Hard Mater. 27
(2009) 688695.
[41] A.L. Prill, H.W. Hayden, J.H. Brophy, Trans. AIME 233 (1965) 19601964.
[42] F. Sailland, J.M. Chaix, C.H. Allibert, Colloquium on Controlling the Properties
of Powder Metallurgy Parts Through Their Microstructures, Societe Francaise
de Metallurgie, Paris, 1990, pp. 4.14.6.
[43] H. Xiong, L. Zhang, Q. Shen, R. Yuan, J. Mater. Sci. Technol. 15 (1999) 9496.
[44] Y. Wu, R.M. German, B. Marx, R. Bollina, M. Bell, Mater. Sci. Eng. A344 (2003)
158167.
[45] P.R. Subramaniam, D.J. Chakrabarti, D.E. Laughlin (Eds.), Monograph Series on
Alloy Phase Diagram-10: Phase Diagramof Binary Copper Alloys, ASMInterna-
tional, Materials Park, OH, 1994, pp. 276286.
[46] J.H. Huang, G.A. Zhou, C.Q. Zhu, S.Q. Zhang, H.Y. Lai, Mater. Lett. 23(1995) 4753.
[47] R.W. Heckel, R.D. Lanam, R.A. Tanzilli, in: J.S. Hirschhorn, K.H. Roll (Eds.),
Advanced Experimental Techniques in Powder Metallurgy, PlenumPress, New
York, 1970, pp. 139188.
[48] C. Padmavathi, Liquid phase sintering of 2712, 6711 and 7775 aluminumalloys
and their properties, Ph.D. Thesis, The Indian Institute of Technology, Kanpur,
2009.
[49] R. Peelamedu, R. Roy, L. Hurtt, D. Agrawal, A.W. Fliet, D. Lewis III, R.W. Bruce,
Mater. Chem. Phys. 88 (2004) 119129.
[50] D.C. Dube, M. Fu, D. Agrawal, R. Roy, A. Santra, Mater. Res. Innovat. 12 (2008)
119122.
[51] J. Cheng, D. Agrawal, Y. Zhang, R. Roy, A.K. Santra, Synthesis of amorphous Si-Ge
alloys using microwave energy, J. Alloys Compd. 491 (2010) 517521.
[52] W. Schatt, E. Friedrich, Powder Metall. 28 (1985) 140144.
[53] A. Upadhyaya, Distortion in liquid phase sintered tungsten heavy alloys, M.S.
Thesis, The Pennsylvania State University, University Park, PA, 1996.
[54] A. Upadhyaya, J.L. Johnson, R.M. German, TungstenandRefractoryMetals-1995,
Metal Powder Industries Federation, Princeton, NJ, 1995, p. 18.
[55] C. Zhou, J. Yi, S. Luo, Y. Peng, L. Li, G. Chen, J. Alloys Compd. 482 (2009) L6L8.
[56] K. Rdiger, K. Dreyer, T. Gerdes, M. Willert-Porada, Int. J. Ref. Met. Hard Mater.
16 (1998) 409416.
[57] V.T. Zaspalis, S. Sklari, M. Kolenbrander, J. Magn. Mater. 310 (2007) 2836.
[58] B.N. Kim, K. Hiraga, K. Morita, H. Yoshida, J. Eur. Ceram. Soc. 29 (2009) 323327.
[59] G.S. Upadhyaya, V. Srikanth, Modern Developments in Powder Metallurgy, vol.
17, Metal Powder Industries Federation, Princeton, NJ, 1985, pp. 5175.
[60] J.L. Chermant, F. Osterstock, J. Mater. Sci. 11 (1976) 19391951.
[61] M.M. Kalina, Z.G. Soboleva, V.N. Gorokhov, E.M. Rabinovich, Powder Metall.
Met. Ceram. 29 (1990) 406409.

You might also like