DR Avijit Mondal - 10 Publication PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 107

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 Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Materials Science and Engineering A 527 (2010) 6870–6878

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Microwave and conventional sintering of 90W–7Ni–3Cu 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 a b s t r a c t

Article history: The present study investigates the possibility of consolidating premixed 90W–7Ni–3Cu alloy – desig-
Received 9 April 2010 nated as 90W–PM (Ni–Cu) – through microwave sintering. An attempt has been made to compare the
Received in revised form 8 June 2010 results between microwave and conventionally sintered samples. This study also compares the sintering
Accepted 22 July 2010
behavior of 90W–7Ni–3Cu with prealloyed 90W–PA (Ni–Cu) in both conventional as well as microwave
furnace at various temperatures. The comparative analysis is based on the sintered density, densifica-
tion parameter, hardness and microstructures of the samples. The present investigation also includes the
Keywords:
variation of matrix composition as a function of temperature by EPMA analysis. The results show that
Microwave sintering
Conventional sintering
microwave sintering requires about 75% less processing time than required by conventional method and
Tungsten heavy alloys still provides better physical and mechanical properties.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction order to obtain good mechanical properties. Ariel et al. [12] corre-
lated the mechanical properties of W–Ni–Cu system with sintered
Tungsten heavy alloy (WHA) is a group of two-phase microstructure. Their study showed that the mechanical properties
composites, based on W–Ni–Cu and W–Ni–Fe alloys. are a function of mean free path between tungsten grains, volume
Tungsten–nickel–copper alloys are widely used for ordinance fraction of tungsten grains and the contiguity of tungsten spheroids.
application, electrical contacts of switches, radiation shielding, The solubility of tungsten in the liquid binder plays a domi-
mass balances, etc. In general the tungsten heavy alloys have been nant role in determining the mechanical properties of the sintered
processed through powder metallurgy route since 1930s [1]. These W–Ni–Cu alloys. The solubility of tungsten in copper is negli-
heavy alloys contain mainly pure tungsten as principal phase in gible. Even at temperatures as high as 1350 ◦ C only 0.04 at.% of
association with a binder phase containing transition metals (Ni, tungsten goes in solution with copper [13]. In contrast, tungsten
Fe, Cu, Co) [2]. In W–Ni–Cu alloys, normally the nickel-to-copper exhibits appreciable solubility (up to 40 wt.%) in nickel. It is there-
ratio ranges from 3:2 to 4:1. Price et al. [3] were the first to propose fore possible to tailor the tungsten solubility, wetting and the
Ni–Cu as the binder for tungsten heavy alloys. Over the last several dihedral angle and thereby, the accompanying sintering response
years, these alloys have been extensively investigated for den- and properties of the system by varying the Ni:Cu ratio [10,14].
sification mechanism, microstructural evolution and properties Nowadays, W–Ni–Cu alloys are also being consolidated by employ-
[4–7]. ing prealloyed powders [15]. Use of prealloyed powder improves
The effect of tungsten and copper powder size variation on homogeneity. The homogenization process accelerates sintering
the sintered properties of W–Ni–Cu heavy alloys was carried out and promotes densification. The alloy formation during sintering
by Srikanth and Upadhyaya [8,9]. The effect of composition and decreases the material viscosity and, hence, stimulates material
sintering temperature on the densification and microstructure of flow under the action of capillary forces [16].
W–Ni–Cu heavy alloys was studied by Ramakrishnan and Upad- Despite widespread application, difficulties still exist in the
hyaya [10]. Kuzmic [11] proposed that rapid cooling from the manufacture of liquid phase sintered tungsten heavy alloys. In
sintering temperature prevents the formation of brittle phase in order to avoid thermal shock, processing of tungsten heavy alloys 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 significant
∗ Corresponding author at: Materials Science and Engineering, Indian Institute of microstructural coarsening during sintering, leading to the degra-
Technology Kanpur, P.O. IIT Kanpur, Kanpur 208016, UP, India. dation of mechanical properties. This problem is further aggravated
Tel.: +91 512 2597672; fax: +91 512 2597505. when the initial powder size is extremely fine. Hence, it is envis-
E-mail address: anishu@iitk.ac.in (A. Upadhyaya).

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) 6870–6878 6871

Table 1
Powder characteristics of as-received elemental W, Ni, Cu and prealloyed Ni–Cu powder.

Property W Ni Cu 69Ni–31Cu

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)


D10 2.0 3.8 19.4 8.2
D50 4.2 11.0 47.0 17.3
D90 6.2 31.8 96.3 40.4

aged that a fast heating rate would mitigate this problem. One of To study the densification behavior; the green (as-pressed) com-
the techniques to achieve fast and relatively uniform sintering is pacts were sintered using conventional and microwave furnace.
through microwaves [17]. The W–Ni–Cu alloys are usually pre- The experimental details have been mentioned elsewhere [32].
pared by mixing the constituent powders. One of the challenges The sintered density was obtained by both dimensional mea-
in the processing of these alloys is to ensure homogenous melt surements as well as Archimedes’ density measurement technique.
distribution [15]. Nickel and copper are known to interdiffuse at To compare the densification response of various compositions, the
temperatures higher than 900 ◦ C. However, in case of rapid sinter- sintered densities were normalized with respect to the theoretical
ing in microwave this may not be readily feasible. density. To take into account the influence of the initial as-pressed
Microwave heating is much more uniform at a rapid rate result- density, the compact sinterability was also expressed in terms of
ing in reduction of processing time and energy consumption. Also densification parameter which is calculated as follows:
rapid heating leads to finer microstructure enhancing the mechan-
ical properties. Clark and Sutton [18] cited many other benefits of sintered density − green density
densification parameter= (1)
the process such as precise and controlled heating, environmen- theoretical density − green density
tally friendly etc. Microwave heating is a very sensitive function An Anter 1161 V vertical dilatometer was used for measuring
of the material being processed and depends on several factors, axial shrinkage and shrinkage rate during sintering with constant
such as sample size, as-pressed density, its mass and geometry heating rate of premixed and prealloyed 90W–7Ni–3Cu compacts.
[19,20]. Though there have been attempts to explain microwave It measures dimensional changes over the entire sintering cycle
heating of metal powders, still there is not yet any consensus on a with precision of 1 ␮m. The dilatometery studies were performed
comprehensive theory to explain the mechanism [21]. at constant heating and cooling rate of 10 ◦ C/min. The sintered
Applicability of microwave sintering to metals was ignored due samples were wet polished in a manual polisher (model: Lunn
to the fact that they reflect microwaves. Roy et al. [22] reported Major, supplier: Struers, Denmark) using a series of 6 ␮m, 3 ␮m and
that particulate metals can be heated rapidly in microwaves. This 1 ␮m diamond paste, followed by cloth polishing using a 0.04 ␮m
has led to the use of microwaves to consolidate a range of partic- colloidal SiO2 suspension. The scanning electron micrographs of as-
ulate metals and alloys [23–25]. While researchers have reported polished samples were obtained by a scanning electron microscope
microstructural refinement due to rapid heating in microwaves, the (model: FEI quanta, Netherlands) in the secondary electron (SE)
effect of microwave sintering on the microstructural homogene- mode. A quantitative analysis was also carried out on selected spec-
ity and mechanical properties seem to be system specific [23–25]. imens using EPMA equipment (model: JXA-8600 SX Super-Probe,
Researchers have also attempted to consolidate refractory materi- supplier: JEOL, Japan). Phase determination and phase evolution, if
als such as pure W, W–Cu, W–Ni–Fe and W–Ni–Cu [26–33] using any were studied for all the samples using an X-ray Diffractometer
microwave energy and reported significant reduction in process (model: Rich. Seifert & Co., GmbH & Co., KG, Germany).
time, elimination of brittle intermetallic formation and superior Bulk hardness measurements were performed on polished sur-
mechanical properties. faces of sintered cylindrical compacts at a load of 5 kg using Vickers
This paper reports the sintering behavior of premixed and pre- hardness tester (model: V100-C1, supplier: Leco, Japan). The load
alloyed 90W–7Ni–3Cu powders in both, microwave as well as was applied for 30 s. Micro hardness tester (model: 8299, supplier:
in a conventional (radiatively heated) furnace. The comparative Leitz, Germany) was used to evaluate the hardness of the matrix
analysis is based on the sintered density, densification parameter, phase. The load applied on matrix phase during the micro hard-
hardness and microstructures of the samples. ness measurement was 15 g. 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
2. Experimental procedure
premixed and prealloyed samples were performed following the
procedure described in MPIF Standard 41 [34].
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 3. Results and discussion
Ni–Cu powder used for this study. For investigating the densifi-
cation response and compaction studies of various compositions, 3.1. Heating response and densification of W–Ni–Cu alloy
cylindrical pellets (diameter: 16 mm and height 8 mm) were
pressed at 200 MPa using a uniaxial semi-automatic hydraulic press Fig. 1 compares the thermal profiles of 90W–7Ni–3Cu compacts
(model: CTM 50, supplier: FIE, India) of 50 T capacity. To facilitate (prepared using premixed and prealloyed Ni–Cu binder) in a con-
compaction and subsequent removal of compacted samples, zinc- ventional and microwave furnace. It is evident from the figure that
stearate powder was applied as die-wall lubricant. The as-pressed W–Ni–Cu alloys couple with microwaves and undergo rapid heat-
green compact density varied from 56% to 58% of the theoretical ing. In case of conventional furnace, in order to ensure uniform
density of the alloy, where the theoretical density was calculated heating, the compacts heating rate was restricted to 5 ◦ C/min and
using the inverse rule of mixtures. isothermal holds were provided at intermittent temperatures. In
Author's personal copy

6872 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 6870–6878

Fig. 2. Photograph of 90W–7Ni–3Cu compacts conventionally sintered at various


temperatures ranging from 1200 ◦ C to 1450 ◦ C.

Fig. 1. Thermal profiles of 90W–7Ni–3Cu alloys heated up to 1450 ◦ C in conven-


tional and microwave furnace. The heating profile of W–Ni–Fe alloys prepared using
premixed and prealloyed Ni–Cu binder is shown separately in case of microwave
processing.

contrast, the overall heating rate achieved in the microwave fur-


nace was 22 ◦ C/min. Taking into consideration the slow heating rate
and isothermal holds associated with conventional sintering, the
compacts were consolidated through microwaves with significant
(∼75%) reduction in the processing time. Figs. 2 and 3 show the pho-
tograph of the 90W–7Ni–3Cu alloys sintered in the temperature
range of 1200–1450 ◦ C in conventional and microwave furnace,
respectively. Note that irrespective of the heating mode and the Fig. 3. Photograph of the 90W–7Ni–3Cu compacts prepared using (a) prealloyed
binder chemistry, none of the 90W–7Ni–3Cu compacts distorted and (b) premixed Ni–Cu binder and sintered in a microwave furnace at various
temperatures.
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 nace. It can therefore be inferred that the chemical state of the
sintered samples. This is attributed to the volumetric heating of alloying constituents influences the microwave coupling and the
the compacts, a unique feature of microwave heating. From the resultant heating. It is well recognized that microwave heating
microwave heating profile (Fig. 1), it is interesting to note that alloys is material dependent. Peelamedu et al. [36] and Roy et al. [37]
with premixed Ni–Cu binder heat at slightly faster rate than those demonstrated in multi-phase systems, the phases heat up at dif-
containing prealloyed matrix. Elsewhere, Sethi et al. [23] and Upad- ferent rates and lead to local an isothermal heating effect. Through
hyaya and Sethi [35] too have reported similar heating response of model experiments in systems, such as Y2 O3 –Fe3 O4 , BaCO3 –Fe3 O4 ,
the premixed and prealloyed Cu–12Sn bronze in microwave fur- and NiO–Al2 O3 , 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) densification parameter of 90W–7Ni–3Cu
alloys. For the sake of comparison, the density and densification 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) 6870–6878 6873

Table 2
Composition as a function of temperature and heating mode of the matrix phase of liquid phase sintered 90W–7Ni–3Cu 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

Conventional 29.25 ± 0.73 53.62 ± 1.00 15.59 ± 0.13


1200
Microwave 27.26 ± 0.96 54.42 ± 0.65 15.20 ± 0.40

Conventional – – –
1250
Microwave 25.12 ± 0.57 55.68 ± 0.78 16.32 ± 0.55

Conventional 26.37 ± 0.38 56.36 ± 0.39 15.36 ± 0.22


1300
Microwave 28.73 ± 18.0 50.34 ± 6.99 17.08 ± 10.63

Conventional 23.25 ± 7.56 56.96 ± 2.3 18.80 ± 5.60


1450
Microwave 19.34 ± 3.09 57.97 ± 2.71 20.62 ± 5.80

species diffuse rapidly into the relatively colder ones and correlated activate densification. Besides enhanced diffusivity at high tem-
it with the very rapid diffusion rates observed in microwave pro- perature, another factor that contributes to densification is the
cessed compacts. It is likely that a similar mechanism is also active solid-solubility of tungsten in the Ni–Cu matrix [38–40]. In one of
in the W heavy alloys system as well. The presence of such effect the very early works, Prill et al. [41] based on the sintering studies of
will result in densification enhancement in the microwave sintered 99W–1(Ni–Cu) alloys showed that the dissolution of the tungsten
components. grains in the matrix and their subsequent reprecipitation from the
To test the above hypothesis, the densification responses of the supersaturated matrix on to the larger tungsten is a major con-
compacts consolidated at various temperatures was evaluated for tributor to densification. To validate this, the effect of sintering
both the heating modes and is summarized in Fig. 4a and b. From temperature, heating mode and the powder type on the matrix
Fig. 4a, it is evident that the density of the compacts increases composition was evaluated using EPMA. The corresponding results
with increasing sintering temperature. Since the green density for 90W–PM (Ni–Cu) and 90W–PA (Ni–Cu) compacts are summa-
of all the compacts were the same, the densification parameter rized in Table 2a and b, respectively. From Table 2a, it is obvious
results (Fig. 4b) closely follow the same trend as sintered den- that higher sintering temperature leads to higher tungsten solubil-
sity. Note that with the exception of 1200 ◦ C, for all other sintering ity in the matrix for the samples prepared from premixed powder.
temperatures, microwave sintering results in better densification. The matrix composition analysis is similar to those reported by Sail-
Since factors that contribute to enhanced diffusivity are known to land et al. [42] and Xiong et al. [43] for the sintered W–Ni–Cu alloys.
enhance densification during sintering, it can be inferred that the On the other hand for the prealloyed compositions (Table 2b) the
atomic mass transport is higher in case of microwave sintering. This tungsten solubility does not follow any specific trend as a function
will also result into less grain coarsening during microwave sinter- of temperature. The complete explanation for this observation is
ing. Consequently, the finer particle size is retained at the sintering not readily available. It is hypothesized that this trend is possibly
temperature which results in higher grain boundary diffusivity. due to the difference in diffusivity of tungsten in prealloyed Ni–Cu
From the sintered density and the densification parameter matrix.
plots shown in Fig. 4, it is evident that for both heating modes
the densification increases with increasing temperature. For the 3.2. Dilatometric evaluation of 90W–7Ni–3Cu alloy
sake of comparison the sintered density and densification param-
eter results reported by Ramakrishnan and Upadhyaya [10] on For better understanding the phenomenology of the densifi-
the similar alloy composition sintered at 1300 ◦ C and 1400 ◦ C has cation of the W–Ni–Cu compacts the dimensional changes were
been superimposed on Fig. 4a and b. Note that the densification measured in situ during sintering using dilatometry. However,
results from this study on 90W–PM (Ni–Cu) are similar to those owing to the instrument’s limitations, dilatometric evaluation
reported by Ramakrishnan and Upadhyaya [10]. Based on the sin- could be performed through conventional heating mode only.
tering studies on a range of W–Ni–Cu alloys, Makipiritti [38] and Fig. 5a and b plots the dilatometric data of shrinkage and shrinkage
Kothari [6] have confirmed that densification in this system is rate vs. temperature for the as-pressed 90W–7Ni–3Cu compacts
diffusion-controlled. Kothari [6] and Crowson [39] have shown prepared using premixed and prealloyed binder. From the figures,
that increasing sintering temperature results in enhanced diffu- it is evident that both compacts undergo gradual densification
sivity and lower activation energy in W–Ni–Cu alloys. The same during heating till 1280 ◦ C. Subsequent heating beyond 1280 ◦ C
will hold good for the present investigation as well and thereby results in sudden increase in the axial shrinkage and shrinkage rate.
Author's personal copy

6874 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 6870–6878

Fig. 5. Dilatometer plots showing the effect of Ni–Cu binder state (premixed vs. pre- Fig. 6. XRD plots showing the effect of heating mode and sintering temperature on
alloyed) on the (a) shrinkage and (b) shrinkage rate of 90W–7Ni–3Cu alloys heated the phase evolution of 90W–7Ni–3Cu alloys prepared using (a) premixed and (b)
up to 1400 ◦ C at a constant heating rate of 10 ◦ C/min in reducing atmosphere. prealloyed Ni–Cu binder.

This spurt in the densification is attributed to the onset of binder homogenous microstructure and lowers the interfacial energy of
melting which results in densification through capillary-induced the tungsten (solid)–matrix (liquid) phase providing a higher den-
rearrangement and solution-reprecipitation. The rapid shrink- sity compact. The analyses of matrix composition of the W–Ni–Cu
age resulting from liquid phase sintering sustains until 1400 ◦ C compacts in the present study (Table 2a and b) – particularly, the
wherein nearly full densification (98%) was achieved. Elsewhere, ones sintered at 1200 ◦ C – are in accordance with this hypothe-
Wu et al. [44] and Martin et al. [15] too have reported simi- sis. At 1200 ◦ C, the dissolved tungsten amount in 90W–PM (Ni–Cu)
lar dilatometric curves for 80W–14Ni–6Cu alloys and prealloyed is lower than its prealloyed counterpart. Consequently, the lat-
90W–6.9Ni–3.1Fe, respectively. For the range of compositions ter exhibits higher densification (Fig. 4b). For compacts heated or
used, both researchers [15,44] have demonstrated that the onset sintered at higher temperatures, the densification in both the com-
of melt formation in the W–Ni–Cu system with Ni and Cu in pacts is not too different and is attributed to homogenization of
the ratio 7:3 occurs at 1282 ◦ C. From the binary phase dia- the premixed matrix through interdiffusion. It is well known that
gram [45] and the experiments [44] the solidus and the liquidus Cu and Ni have mutual intersolubility [45]. Using X-ray analysis,
temperatures of the 70Ni–30Cu matrix was determined to be Heckel et al. [47] have experimentally shown that significant inter-
1345 ◦ C and 1380 ◦ C, respectively. However, in case of ternary diffusion and homogenization through in situ alloying occurs in the
W–Ni–Cu system, due to substantial solubility of the tungsten Ni–52 at.% Cu compact after sintering at 950 ◦ C for 1 h. Since the
in the matrix the solidus and liquidus temperatures are lowered Cu:Ni ratio in the present system (90W–7Ni–3Cu) is much lower,
[42]. hence, it is quite reasonable to assume that the premixed Ni–Cu
From the dilatometric results (Fig. 5a and b), it can be dis- powders will get completely homogenized during conventional
cerned that the 90W–PA (Ni–Cu) alloy exhibits slightly higher sintering.
densification particularly at lower temperatures as compared to the
90W–PM (Ni–Cu) compact. This trend is similar to that reported 3.3. Microstructural evolution and properties
by Huang et al. [46] who investigated the influence of preal-
loyed Ni–Fe–Mo binder on the sintering behavior of tungsten Fig. 6a and b compares the effect of heating mode and
heavy alloys. They [46] noted that prealloying results in a more sintering temperature on the phase evolution in 90W–PM
Author's personal copy

A. Mondal et al. / Materials Science and Engineering A 527 (2010) 6870–6878 6875

Fig. 7. SEM micrographs of 90W–7Ni–3Cu alloys prepared using prealloyed Ni–Cu binder and consolidated in a microwave furnace at (a) 1200 ◦ C, (b) 1250 ◦ C, (c) 1300 ◦ C
and (d) 1450 ◦ C, respectively.

(Ni–Cu) and 90W–PA (Ni–Cu) alloys, respectively. For all [48] in case of microwave sintering of various aluminium
XRD experiments, the sample size and other test parame- alloys.
ters were kept constant. It is therefore rather surprising to The reduction in the intensity indicates that the microwave
note that in case of microwave sintered compacts some of sintered compacts have less perfect crystals and contain more
the peaks do not even register on the diffractogram. Sim- defects as compared to their conventionally sintered counter-
ilar observations have also been reported by Padmavathi parts. Recently, many researchers, [37,49–51] have reported that

Fig. 8. SEM micrographs of (a) premixed and (b) prealloyed 90W–7Ni–3Cu 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) 6870–6878

Fig. 9. SEM micrographs of (a) premixed and (b) prealloyed 90W–7Ni–3Cu alloys sintered at 1450 ◦ C in conventional (left) and microwave (right) furnace.

exposure to microwave heating leads to decrystallization and the prealloyed sintered compacts have slightly higher coarsening
sometimes even amorphization in some of the ceramic and semi- than their premixed counterparts. This can be attributed to a more
conducting systems. The defects within a crystalline structure are homogenous microstructure and better spreading of the melt due
known to activate sintering [52]. Hence, it may be possible that to the relatively lower tungsten (solid)–matrix (liquid) interfacial
in the present system too the defects generated on exposure to energy in case of alloys prepared using prealloyed binder [46]. More
microwaves contribute to densification enhancement. recently, Martin et al. [15] showed that the Ni–Cu additive phase
The phenomenology of microstructural evolution in conven- in 90W–6.9Ni–3.1Cu alloy with prealloyed matrix is more effec-
tionally sintered W–Ni–Cu alloys has been extensively investigated tively distributed within the tungsten grains during sintering and
by several researchers [3,10,15,40]. In the present study, the atten- contributes not only to densification but also leads to enhanced
tion was focused on the microstructural response of microwave tungsten grain coarsening as compared to premixed alloys.
sintered W–Ni–Cu system and its comparison with that obtained One of the intriguing observations in the present study is the
through conventional sintering. Fig. 7a–d shows the effect of sinter- dramatic difference between the microstructure of W–Ni–Cu com-
ing temperature on the representative microstructure of 90W–PA pacts sintered in conventional and microwave furnace (Fig. 8a and
(Ni–Cu) alloy consolidated in a 2.45 GHz multimode microwave fur- b). In case of conventionally sintered compacts, the microstructure
nace. It is quite evident that higher temperature not only results in at 1300 ◦ C is quite similar to that obtained through solid-state sin-
pore elimination but is also accompanied by redistribution of the tering. Some of the earlier researchers [10,53,54] too have reported
matrix and coarsening of tungsten grains. It is worth noting that similar findings at this temperature for premixed 90W–7Ni–3Cu
the microstructure of the compacts sintered in solid-state (1200 ◦ C compacts. The larger tungsten grain size in microwave sintered
and 1250 ◦ C) is drastically different from the liquid phase sintered W–Ni–Cu compacts at 1300 ◦ C is contrary to the observations in
counterparts (Fig. 7c and d). In case of compacts sintered at 1200 ◦ C, various systems that a fast heating rate will result in lower grain size
the binder phase is inhomogenously distributed (Fig. 7a) and the [30,55,56]. However, similar contrary results have recently been
tungsten grains are faceted. In contrast, at 1250 ◦ C (Fig. 7b), while reported by Zaspalis et al. [57] in microwave sintered Mn–Zn fer-
the W-grains still remain faceted, the binder is more uniformly dis- rites and by Kim et al. [58] in spark plasma sintered alumina. The
tributed. The well rounded and coarser grains associated with liquid exact explanation of such observations is still lacking and further
phase sintering (Fig. 7c and d) are associated with higher solid- detailed experiments need to be carried out in future to elucidate
solubility (Table 2) and relatively faster solution-reprecipitation this better. In contrast to 1300 ◦ C sintered compacts, both con-
kinetics. ventional as well as microwave sintered 90W–7Ni–3Cu alloys (at
Fig. 8a and b compares the effect of heating mode on the 1450 ◦ C) exhibit well-defined liquid phase sintered microstructure
microstructures of the 90W–PM (Ni–Cu) and 90W–PA (Ni–Cu) with rounded tungsten grains interspersed in the Ni–Cu matrix
alloys, respectively, sintered at 1300 ◦ C. The corresponding (Fig. 9a and b). Table 3 summarizes the effect of heating mode
microstructures of compacts liquid phase sintered at 1450 ◦ C are on the average tungsten grain size in premixed and prealloyed
shown in Fig. 9a and b. From Figs. 8 and 9, it can be discerned that 90W–7Ni–3Cu alloys liquid phase sintered at 1450 ◦ C. From Fig. 9
irrespective of the sintering temperature and the heating mode, 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) 6870–6878 6877

Table 3
Effect of heating mode on the average tungsten grain size in 90W–7Ni–3Cu alloys
prepared using premixed and prealloyed Ni–Cu binder and sintered at 1450 ◦ C.

Composition Heating mode W grain size (␮m)

90W–PM Conventional 23 ± 10
(Ni–Cu) Microwave 21 ± 7

90W–PA Conventional 26 ± 8
(Ni–Cu) Microwave 20 ± 7

Table 4
Effect of heating mode and sintering temperature on the bulk hardness (in kgf/mm2 )
of 90W–7Ni–3Cu alloys prepared using premixed and prealloyed binder.

Heating mode 90W–PM (7Ni–3Cu)a 90W–PA (7Ni–3Cu)

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 composition has been listed as 175VHN and 220VHN for compacts sintered
at 1300 ◦ C and 1400 ◦ C, respectively.
Fig. 10. Effect of heating mode and sintering temperature on the transverse rupture
strength of 90W–7Ni–3Cu alloys prepared using premixed and prealloyed binders.

tial binder chemistry (premixed vs. prealloyed), the microwave


sintered compacts have relatively more refined microstructure Fig. 10 shows the effect of heating mode and powder chem-
as compared to those consolidated in a conventional (radiatively istry on the transverse rupture strength of the 90W–7Ni–3Cu alloy
heated) furnace. Elsewhere, Upadhyaya et al. [30] has reported sintered at 1300 ◦ C and 1450 ◦ C. One of the very interesting observa-
similar findings in liquid phase sintered W–Ni–Fe system. tions is that microwave sintered compacts have higher transverse
The differences in the sintered microstructure of alloys consol- rupture strength as compared to those sintered conventionally. The
idated through conventional and microwave furnace is bound to enhancement of strength is as high as up to threefolds (∼211%) in
influence the mechanical properties. Table 4 summarizes the effect case of alloys processed at 1300 ◦ C. This can be correlated to a less
of heating mode and powder chemistry (prealloyed vs. premixed) contiguous microstructure consisting of rounded tungsten grains
on the bulk hardness of 90W–7Ni–3Cu alloys sintered at 1300 ◦ C that result in alloys microwave sintered at 1300 ◦ C (Fig. 8). Else-
and 1450 ◦ C, respectively. To provide a basis of comparison, the where, Chermant and Osterstock [60] have demonstrated similar
bulk hardness of premixed 90W–7Ni–3Cu alloy sintered at 1300 ◦ C influence of microstructure on the fracture toughness of WC–Co
and 1400 ◦ C reported by Ramakrishnan and Upadhyaya [10] is also system. The transverse rupture strength of the conventionally sin-
provided in Table 4. It is remarkable to note that the hardness tered 90W–PM (Ni–Cu) system is similar to that reported earlier
achieved in the present study is far superior to those reported in by Kalina et al. [61]. Another observation that can be inferred from
earlier work [10,59]. From Table 4, it is worth noting that for all Fig. 10 is that the transverse rupture strength of alloys prepared
conditions, microwave sintering results in up to 13% improvement with prealloyed Ni–Cu is marginally higher than those prepared
in the bulk hardness. This can be attributed to the increased densi- using premixed binder. This could be due to better homogeneity of
fication and a more refined microstructure in case of microwave the alloying additives in case of the prealloyed system.
sintered alloys. Previous report from Upadhyaya et al. [30] has
shown that the hardness of tungsten grains per se does depend 4. Conclusions
on the heating mode. However, on account of the varying tungsten
solid-solubility, the hardness of the matrix phase is expected to 90W–7Ni–3Cu alloys, compacts were prepared using both
be dependent upon the heating mode. To test this, Table 5 summa- premixed and prealloyed Ni–Cu binder and were sintered at
rizes the corresponding matrix hardness measurements performed temperatures ranging between 1200 ◦ C and 1450 ◦ C. As com-
on compacts consolidated in both the heating modes at 1300 ◦ C and pared to conventional sintering, both premixed and prealloyed
1450 ◦ C. It is interesting to observe that in case of compacts con- 90W–7Ni–3Cu alloys were microwave sintered in significantly
solidated at 1300 ◦ C, the matrix hardness is higher for microwave (∼75%) less time. In spite of higher heating rate in microwave sin-
processed compacts (Table 5). This can be correlated to the higher tering, no micro- or macro-cracking was observed in all microwave
tungsten solubility observed in microwave sintered alloys (Table 2a sintered samples. Dilatometric analysis of the 90W–7Ni–3Cu alloy
and b). In contrast, the trend is reversed for compacts sintered at at slower heating rates (<10 ◦ C/min) reveals that – unlike the W–Cu
1450 ◦ C. In case of prealloyed compacts, this can be correlated to system – the W–Ni–Cu system exhibits significant solubility of
the higher W-solubility in conventionally sintered alloys. However, tungsten in the matrix. Consequently, a major portion of the overall
further tests need to be performed to gain detailed insight for this. densification 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 form a homogeneous
Table 5 solid solution. Hence, the sintering response of the 90W–7Ni–3Cu
Effect of heating mode on the micro hardness (in kgf/mm2 ) of the matrix phase in alloy was per se independent of the initial matrix type (premixed
sintered 90W–7Ni–3Cu alloys. vs. prealloyed). An interesting finding in the present study was the
Heating mode 90W–PM (7Ni–3Cu) 90W–PA (7Ni–3Cu) observation of the heating mode effect on the microstructure of the
◦ ◦
compacts sintered at 1300 ◦ C. While the conventionally processed
1300 C 1450 C 1300 ◦ C 1450 ◦ C
alloy had a typical solid-state sintered structure, the microwave
Conventional 260 ± 20 333 ± 13 250 ± 25 320 ± 16 processed samples had a well-developed microstructure, typically
Microwave 290 ± 23 312 ± 15 280 ± 21 290 ± 12
associated with liquid phase sintering. Consequently, as compared
Author's personal copy

6878 A. Mondal et al. / Materials Science and Engineering A 527 (2010) 6870–6878

to conventional sintering, microwave sintering resulted in about [29] A. Mondal, A. Upadhyaya, D. Agrawal, J. Microw. Power Electromagn. Energy
9% improvement in hardness (to ∼350VHN) and about threefold (JMPEE) 43 (2009) 11–16.
[30] A. Upadhyaya, S.K. Tiwari, P. Mishra, Scripta Mater. 56 (2007) 5–8.
(∼211%) enhancement in flexural strength to around 1625 MPa. [31] A. Mondal, D. Agrawal, A. Upadhyaya, J. Microw. Power Electromagn. Energy
(JMPEE) 44 (2010).
Acknowledgement [32] A. Mondal, D. Agrawal, A. Upadhyaya, Mater. Sci. Technol. (2008) 2502–2515.
[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
This collaborative research was done as a part of the Center for of Technology, Kanpur, 2009.
Development of Metal–Ceramic Composites through Microwave [34] MPIF Standard 41, Determination of transverse rupture strength of powder
metallurgy materials, standard test methods for metal powders and powder
Processing which was funded by the Indo-US Science and Tech- metallurgy products, MPIF, Princeton, NJ, USA, 2002, pp. 55–57.
nology Forum (IUSSTF), New Delhi. [35] A. Upadhyaya, G. Sethi, Scripta Mater. 56 (2007) 469–472.
[36] R.D. Peelamedu, R. Roy, D. Agrawal, Mater. Res. Bull. 36 (2001) 2723–2739.
[37] R. Roy, Y. Fang, J. Cheng, D.K. Agrawal, J. Am. Ceram. Soc. 88 (2005) 1640–1642.
References [38] S. Makipiritti, Powder Metall. 4 (1961) 97–111.
[39] A. Crowson, Modern Developments in Powder Metallurgy, vol. 15, Metal Pow-
[1] S.G. Caldwell, in: P.W. Lee, R. Lacocca (Eds.), Powder Metal Technologies and der Industries Federation, Princeton, NJ, 1985, pp. 507–520.
Applications, vol. 7, Materials Park, OH, USA, 1998, pp. 914–921. [40] G. Sethi, S.J. Park, J.L. Johnson, R.M. German, Int. J. Ref. Met. Hard Mater. 27
[2] A. Upadhyaya, Mater. Chem. Phys. 67 (2001) 101–110. (2009) 688–695.
[3] G.H.S. Price, C.J. Smithells, S.V. Williams, J. Inst. Met. 62 (1938) 239–264. [41] A.L. Prill, H.W. Hayden, J.H. Brophy, Trans. AIME 233 (1965) 1960–1964.
[4] A.L. Prill, Trans. AIME 230 (1964) 769–772. [42] F. Sailland, J.M. Chaix, C.H. Allibert, Colloquium on Controlling the Properties
[5] J.H. Brophy, H.W. Hayden, J. Wulff, Trans. TMS-AIME 224 (1962) 797–803. of Powder Metallurgy Parts Through Their Microstructures, Societe Francaise
[6] N.C. Kothari, J. Less Common Met. 13 (1967) 457–468. de Metallurgie, Paris, 1990, pp. 4.1–4.6.
[7] R. Makarov, O.K. Teodorovich, I.N. Fruntsevich, Sov. Powder Metall. Met. Ceram. [43] H. Xiong, L. Zhang, Q. Shen, R. Yuan, J. Mater. Sci. Technol. 15 (1999) 94–96.
4 (1965) 554–559. [44] Y. Wu, R.M. German, B. Marx, R. Bollina, M. Bell, Mater. Sci. Eng. A344 (2003)
[8] V. Srikanth, G.S. Upadhyaya, Int. J. Ref. Hard Met. 2 (1983) 49–54. 158–167.
[9] V. Srikanth, G.S. Upadhyaya, Powder Technol. 39 (1984) 61–65. [45] P.R. Subramaniam, D.J. Chakrabarti, D.E. Laughlin (Eds.), Monograph Series on
[10] K.N. Ramakrishnan, G.S. Upadhyaya, J. Mater. Sci. Lett. 9 (1990) 456–459. Alloy Phase Diagram-10: Phase Diagram of Binary Copper Alloys, ASM Interna-
[11] J.F. Kuzmic, in: H.H. Hausner (Ed.), Modern Development in Powder Metallurgy, tional, Materials Park, OH, 1994, pp. 276–286.
vol. 3, Plenum Press, New York, 1966, pp. 166–171. [46] J.H. Huang, G.A. Zhou, C.Q. Zhu, S.Q. Zhang, H.Y. Lai, Mater. Lett. 23 (1995) 47–53.
[12] E. Ariel, J. Batra, D. Brandon, Powder Metall. Int. 5 (1973) 125–128. [47] R.W. Heckel, R.D. Lanam, R.A. Tanzilli, in: J.S. Hirschhorn, K.H. Roll (Eds.),
[13] V.N. Eremenko, R.V. Minakova, M.M. Churakov, Sov. Powder Metall. Met. Ceram. Advanced Experimental Techniques in Powder Metallurgy, Plenum Press, New
15 (1976) 283–286. York, 1970, pp. 139–188.
[14] Y.V. Naidich, I.A. Lavrinenko, V.A. Evdokimov, Sov. Powder Metall. Met. Ceram. [48] C. Padmavathi, Liquid phase sintering of 2712, 6711 and 7775 aluminum alloys
16 (1977) 276–280. and their properties, Ph.D. Thesis, The Indian Institute of Technology, Kanpur,
[15] J.M. Martin, J.L. Johnson, R.M. German, F. Castro, in: E. Daver, C.J. Trombino 2009.
(Eds.), Advances in Powder Metallurgy and Particulate Materials, vol. 1, Metal [49] R. Peelamedu, R. Roy, L. Hurtt, D. Agrawal, A.W. Fliflet, D. Lewis III, R.W. Bruce,
Powder Industries Federation, Princeton, NJ, 2006, pp. 5.43–5.57. Mater. Chem. Phys. 88 (2004) 119–129.
[16] A.P. Savitskii, Sci. Sinter. 37 (2005) 3–17. [50] D.C. Dube, M. Fu, D. Agrawal, R. Roy, A. Santra, Mater. Res. Innovat. 12 (2008)
[17] K.J. Rao, P.D. Ramesh, Mater. Bull. Sci. 18 (1995) 447–465. 119–122.
[18] D.E. Clark, W.H. Sutton, Ann. Rev. Mater. Sci. 26 (1996) 299–331. [51] J. Cheng, D. Agrawal, Y. Zhang, R. Roy, A.K. Santra, Synthesis of amorphous Si-Ge
[19] R.M. Hutcheon, M.S. De Jong, F.P. Adams, Microw.: Theor. Appl. Mater. Proc. III, alloys using microwave energy, J. Alloys Compd. 491 (2010) 517–521.
Ceram. Trans., Am. Ceram. Soc. 59 (1995) 215. [52] W. Schatt, E. Friedrich, Powder Metall. 28 (1985) 140–144.
[20] A. Mondal, D. Agrawal, A. Upadhyaya, J. Microw. Power. Electromagn. Energy [53] A. Upadhyaya, Distortion in liquid phase sintered tungsten heavy alloys, M.S.
(JMPEE) 43 (2009) 5–10. Thesis, The Pennsylvania State University, University Park, PA, 1996.
[21] M.W. Porada, Microw.: Theor. Appl. Mater. Proc. III, Ceram. Trans., Am. Ceram. [54] A. Upadhyaya, J.L. Johnson, R.M. German, Tungsten and Refractory Metals-1995,
Soc. 80 (1997) 153. Metal Powder Industries Federation, Princeton, NJ, 1995, p. 18.
[22] R. Roy, D.K. Agrawal, J.P. Cheng, S. Gedevanishvili, Nature 399 (1999) 668–670. [55] C. Zhou, J. Yi, S. Luo, Y. Peng, L. Li, G. Chen, J. Alloys Compd. 482 (2009) L6–L8.
[23] G. Sethi, A. Upadhyaya, D. Agrawal, Sci. Sinter. 35 (2003) 49–65. [56] K. Rödiger, K. Dreyer, T. Gerdes, M. Willert-Porada, Int. J. Ref. Met. Hard Mater.
[24] R.M. Anklekar, D.K. Agrawal, R. Roy, Powder Metall. 44 (2001) 355–362. 16 (1998) 409–416.
[25] K. Saitou, Scripta Mater. 54 (2006) 875–879. [57] V.T. Zaspalis, S. Sklari, M. Kolenbrander, J. Magn. Mater. 310 (2007) 28–36.
[26] M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding, D. Kapoor, D. Agar- [58] B.N. Kim, K. Hiraga, K. Morita, H. Yoshida, J. Eur. Ceram. Soc. 29 (2009) 323–327.
wal, J. Chang, Int. J. Powder Metall. 42 (2006) 45–50. [59] G.S. Upadhyaya, V. Srikanth, Modern Developments in Powder Metallurgy, vol.
[27] G. Prabhu, A. Charkraborty, B. Sarma, Int. J. Ref. Met. Hard Mater. 27 (2009) 17, Metal Powder Industries Federation, Princeton, NJ, 1985, pp. 51–75.
545–548. [60] J.L. Chermant, F. Osterstock, J. Mater. Sci. 11 (1976) 1939–1951.
[28] S.D. Luo, J.H. Yi, Y.L. Guo, Y.D. Peng, L.Y. Li, J.M. Ran, J. Alloys Compd. 473 (2009) [61] M.M. Kalina, Z.G. Soboleva, V.N. Gorokhov, E.M. Rabinovich, Powder Metall.
L5–L9. Met. Ceram. 29 (1990) 406–409.
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 Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Journal of Alloys and Compounds 509 (2011) 301–310

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Effect of heating mode and sintering temperature on the consolidation of


90W–7Ni–3Fe alloys
Avijit Mondal a , Anish Upadhyaya b,∗ , Dinesh Agrawal c
a
NTPC Energy Technology Research Alliance (NETRA), National Thermal Power Corporation Ltd., Greater Noida 201308, India
b
Department of Materials Science and Engineering Indian Institute of Technology, Kanpur 208016, India
c
The Pennsylvania State University, University Park, PA 16802, USA

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

Article history: The present study compares the sintering response of 90W–7Ni–3Fe alloys consolidated in a 2.45 GHz
Received 6 April 2010 microwave furnace and a conventional furnace. The W–Ni–Fe compacts were sintered in a temperature
Received in revised form 28 August 2010 range of 1200–1500 ◦ C corresponding to solid-state as well as liquid phase sintering. The compacts were
Accepted 2 September 2010
successfully sintered in a microwave furnace with about 80% reduction in the overall processing time.
Available online 21 September 2010
For both the heating modes, the W–Ni–Fe alloys exhibited significant densification prior to melt for-
mation through solid-state sintering. The in situ dilatometric studies revealed that the contribution to
Keywords:
densification from solid-state sintering is higher at lower heating rates. In comparison to conventional
Sintering
Microwave heating
sintering, microwave sintered compacts showed relatively refined microstructure and higher hardness
Tungsten heavy alloy and flexural strength.
Microstructure © 2010 Elsevier B.V. All rights reserved.

1. Introduction and found that ductility of the alloys continuously decreases


with increasing tungsten content and an optimal combination of
Tungsten heavy alloys (WHAs) belong to a group of two-phase strength and ductility is achieved at about 93 wt.% tungsten. Bel-
composites, based on W–Ni–Cu and W–Ni–Fe alloys. Although hadjhamida and German [8] too confirmed the decrease in ductility
WHAs with Ni–Cu binder were initially developed, attention has with increasing tungsten content. However, they found that the
now been focussed on heavy alloys with Ni–Fe binder because of strength of W–Ni–Fe alloy increases with an increasing tungsten
their superior mechanical properties [1–3]. WHAs possess unique content and attains a maximum at about 85 vol.% (93 wt.%) tung-
combination of properties, such as high density (16–18 g/cm3 ), high sten. Besides tungsten powder size and distribution, the properties
strength (1000–1700 MPa) and high ductility (10–30%) [3], ther- of W–Ni–Fe alloys are also influenced by sintering time and tem-
mal conductivity and corrosion resistance. This makes them unique perature. Bourguignon and German [9] reported lower strength
in many applications such as radiation shields, vibration dampers, and higher ductility of 93W–4.9Ni–2.1Fe alloy liquid phase sin-
kinetic energy penetrators and heavy-duty electrical contacts. tered at higher temperature. The strength decreased from 924 MPa
The highly dense W–Ni–Fe heavy alloys could be fabricated to 899 MPa and the ductility increased from 14% to 27% when the
only by liquid phase sintering [4]. Dzykovich et al. [5] have quan- sintering temperature was increased from 1465 ◦ C to 1580 ◦ C. This
titatively analyzed the distribution of individual elements in the was attributed to the rate of tungsten grain growth, which results
tungsten grains and matrix phase in sintered W–Ni–Fe alloys. They in proportionately higher fraction of matrix phase between the
observed that addition of Fe decreases the solubility of W in Ni grains. In yet another study, Bose et al. [10] reported similar effect
and prevents the formation of an intermetallic phase. To avoid this, of temperature on strength and ductility of liquid phase sintered
the matrix composition is restricted to an optimal nickel to iron 90W–7Ni–3Fe system.
ratio of 7:3 [5,6]. The sintered properties of WHAs exhibit a com- German et al. [11] studied the effect of sintering time on the
plex dependence on particle size, matrix composition, sintering microstructure and mechanical properties of tungsten heavy alloys.
time, temperature and atmosphere, heating/cooling rate, and post- They reported that increasing sintering time obviously resulted in
sintering heat-treatment. Rabin et al. [7] correlated the strength microstructural coarsening, which degraded the mechanical prop-
and ductility of W–Ni–Fe heavy alloys with the microstructure erties. During the heating process, substantial changes in terms
of thermodynamics and microstructure occur with the tempera-
ture to minimize the free energy. Clearly, the changes are time and
∗ Corresponding author. Tel.: +91 512 2597672; fax: +91 512 2597505. heating rate dependent. Recently, Bollina and German [12] have
E-mail address: anishu@iitk.ac.in (A. Upadhyaya). critically examined the effect of heating rate in conventional tech-

0925-8388/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jallcom.2010.09.008
Author's personal copy

302 A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310

Fig. 1. SEM micrographs of as-received (a) W, (b) Ni, and (c) Fe powder.

nique in a range from 1 ◦ C/min to 15 ◦ C/min, on densification during was carried out in a MoSi2 -heated horizontal tubular furnace (model: OKAY 70T-7,
liquid phase sintering of a range of W–Ni–Fe alloy compositions. supplier: Bysakh, Kolkata, India) at a constant heating rate of 5 ◦ C/min. The sinter-
ing temperatures selected for the solid and liquid phase sintering of 90W–7Ni–3Fe
Conventional sintering requires long soaking time for W–Ni–Fe
alloys were 1200 ◦ C, 1250 ◦ C, 1300 ◦ C, 1450 ◦ C and 1500 ◦ C, and sintering time was
alloys which results in undesired grain growth and precipitation of kept constant at 60 min. All samples were processed in flowing H2 atmosphere
brittle intermetallic phases. Microwave sintering of W–Ni–Fe com- with dew point −35 ◦ C. Microwave sintering of the green compacts was carried out
positions was studied by Upadhyaya et al. [13]. They investigated using a multi-mode cavity, 2.45 GHz, 6 kW commercial microwave furnace (model:
the sintering response of a 92.5W–7.5(Ni–Fe) alloy – with a non- RC/20SE, supplier: Amana Radarange). Further details of the microwave furnace and
experimental arrangements have been described elsewhere [17].
optimal matrix composition – consolidated through microwaves The sintered density was obtained by both dimensional measurements as well
and compared its densification, microstructure and mechanical as Archimedes principle. In order to take into account the influence of the ini-
properties with those of conventionally sintered compacts. Many tial as-pressed density, the compact sinterability was also expressed in terms of
studies on microwave sintering of tungsten and its alloys have densification parameter which is calculated as follows:
been conducted earlier by several researchers [14–23]. Zhou et (Sintered density − green density)
Densification parameter = (2)
al. [20] investigated the effect of heating rate on the microwave (Theoretical density − green density)
sintered 90W–7Ni–3Fe heavy alloys, with the heating rate in the To compare the densification response of various compositions, the sintered
range from 10 ◦ C/min to 112 ◦ C/min. They have shown that a faster densities were normalized with respect to the theoretical density. The formulae
heating rate results in smaller W grain size and larger W–W con-
tiguity. The heating rate of 80 ◦ C/min has the best combination of
the microstructure and mechanical performance. The present study
evaluates the effect of heating mode and sintering temperature on
the sintering of 90W–7Ni–3Fe alloys.

2. Experimental procedure

The as-received powders were characterized for their size, size distribution
and morphology. Table 1 summarizes the characteristics of the as-received W, Ni,
Fe powders used in the present study. Fig. 1 shows the scanning electron micro-
graphs of the precursor powders. The chemically reduced tungsten powder has
irregular, faceted morphology. Nickel powder, which was produced by carbonyl
decomposition, resulted in a rounded morphology with spiky appearance. Iron pow-
der was also produced by carbonyl decomposition. The alloy compositions were
prepared by mixing the required proportions of each powder in a turbula mixer
(model: T2C Nr 921266, supplier: Bachofen AG, Switzerland) for 60 min to ensure
complete homogeneity in the resultant powder mixture. For investigating the densi-
fication response and compaction studies of various compositions, cylindrical pellets
(diameter: 1.6 cm and height 0.8 cm) were pressed at 200 MPa using a uniaxial, semi-
automatic hydraulic press (model: CTM 50, supplier: FIE, Ichalkaranji, India) and
zinc-stearate powder as die-wall lubricant. The as-pressed green compact density
varied from 54% to 56% of the theoretical density of the alloy, calculated using the
inverse rule of mixtures.
To study the densification behavior, the green compacts were sintered using Fig. 2. Thermal profile of 90W–7Ni–3Fe alloy consolidated in a radiatively heated
conventional and microwave furnace. The conventional sintering of green compacts (conventional) and microwave furnace.
Author's personal copy

A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310 303

Table 1
Characteristics of the as-received powders used in the present study.

Powders W Ni Fe

Supplier Kennametal Widia India Ltd. INCO Kennametal Widia India Ltd.
Processing route Chemical reduction Carbonyl process Carbonyl process
Powder shape Irregular Spiky Spherical
Flow rate (s/50 g) Non-flowing Non-flowing 5.5

(Powder size, ␮m)


D10 2.0 3.8 1.4
D50 4.2 11.0 7.0
D90 6.2 31.8 19.0

Apparent density, g/cm3 (%theoretical) 4.8 (25%) 2.2 (25%) 3.1 (39%)
Tap density, g/cm3 (%theoretical) 6.3 (33%) 3.1 (35%) 3.7 (47%)

used for calculation of axial (ıA ) and radial shrinkage (ıR ) of the sintered cylindrical are similar to those reported by Upadhyaya et al. [13] in case of
compacts are given below: 92.5W–6.4Ni–1.1Fe alloys and more recently by Zhou et al. [20] in
  case of 90W–7Ni–3Fe system. Fig. 3 compares the temporal power
hs
ıA = 1− × 100% (3)
hg consumption in heating the 90W–7Ni–3Fe alloys. It is interesting to
  note that heating of the same compact in microwaves is achieved at
rs
ıR = 1− × 100% (4) much lower power consumption. This can be attributed to the fact
rg
that in microwaves only the sample per se heats up and acts as the
where hg and hs are heights; and rg and rs are radii of the green and sintered compact, source of heat. As a result, energy is primarily consumed in heat-
respectively.
ing up the sample only. Consequently, the effective thermal mass
For understanding the phenomenology of phase evolution, differential scanning
calorimetry (DSC) was used. The W–Ni–Fe compacts were heated to 1500 ◦ C at a reduction lowers the required power input. Despite such high heat-
constant rate of 10 ◦ C/min and subsequently cooled at a controlled rate (10 ◦ C/min) ing rate, no micro- or macro-cracking and distortion was observed
in a DSC unit (model: SDT2960, supplier: TA Instruments, New Castle, DE, USA). in microwave sintered samples (Fig. 4). This further underscores the
For measuring in situ axial shrinkage and shrinkage rate during sintering efficacy of unique volumetric heating associated with microwaves.
dilatometry experiments were performed using a vertical push rod dilatometer
in reducing atmosphere. The temperature inside the dilatometer (model: 1161V,
Anter, Pittsburgh, PA, USA) was accurate to ±2 ◦ C. It measures dimensional changes 3.2. Effect of heating mode and sintering temperature on
over the entire sintering cycle with precision of 1 ␮m. The dilatometry studies were densification
conducted at various heating rates (2 ◦ C/min, 5 ◦ C/min, and 10 ◦ C/min) and constant
cooling rate of 10 ◦ C/min. Metallographic techniques were employed on the sintered
samples. The sintered samples were wet polished in a manual polisher (model: Lunn Fig. 5a and b show the effect of sintering temperature on the
Major, supplier: Struers, Denmark) using a series of 6 ␮m, 3 ␮m and 1 ␮m diamond density and densification parameter of 90W–7Ni–3Fe alloy con-
paste, followed by cloth polishing using a suspension of 0.04 ␮m colloidal SiO2 sus- solidated in conventional and microwave furnaces, respectively.
pension. The scanning electron micrographs of as-polished samples were obtained
In case of conventional sintering the compacts were sintered up
by a scanning electron microscope (model: Quanta, supplier: FEI, The Netherlands)
in the secondary electron (SE) mode. Grain size measurement was done by line to 1500 ◦ C for 1 h. However, owing to slight distortion that was
intercept method. During the measurement it was assumed that there were no sub discerned in microwave sintering the compacts sintering tempera-
grain boundaries inside the tungsten grain. In the present study, quantitative anal- ture in microwave furnace was restricted to 1450 ◦ C, 30 min. In fact,
ysis was carried out on selected specimens using an EPMA (model: JXA-8600 SX microwave sintered compacts attained nearly full (>98.5%) density
Super-Probe, supplier: JEOL, Japan). Phase determination and phase evolution, if any
were studied for all the samples using X-Ray Diffractometer (model: Rich. Seifert
at 1450 ◦ C, 30 min. Prolonged soaking time in microwave led to
& Co., GmbH & Co. KG, Germany). Bulk hardness measurements were performed distortion of the compact and therefore was not employed.
on polished surfaces of sintered cylindrical compacts at a load of 5 kg using Vick- For both the heating modes, the sintered density increases with
ers hardness tester (model: V100-C1, supplier: Leco, Japan). The load was applied sintering temperature. Since the green density of the compacts was
for 30 s. Micro hardness tester (model: 8299, supplier: Leitz, Germany) was used to
evaluate the hardness of tungsten grains and the matrix phase. The loads applied on
tungsten grain and matrix phase during the micro hardness measurement were 50 g
and 15 g, respectively. The bulk hardness and the micro hardness values of tungsten
grain and matrix phase reported for each sample were an average of five read-
ings taken at different locations on respective phases. Transverse rupture strength
(TRS) measurements of sintered samples were performed following the procedure
described in MPIF Standard 41 [24].

3. Results and discussion

3.1. Heating response of W–Ni–Fe alloy

Fig. 2 compares the thermal profiles for 90W–7Ni–3Fe alloy con-


solidated in conventional and microwave furnaces, respectively.
From the figure, it is quite evident that 90W–7Ni–3Fe samples
effectively couple with the microwaves and undergo rapid heating.
The overall heating rate observed in case of microwave heating
was 29 ◦ C/min. It is also interesting to note that due to the less
thermal mass, the cooling rate in microwave furnace is relatively
higher. Taking into consideration the slow heating rate (5 ◦ C/min)
and isothermal holds at intermittent temperatures in conventional
furnace, microwave sintering results in about 80% reduction in Fig. 3. Input power variation with time in heating of 90W–7Ni–3Fe alloy compacts
the overall process time. These scales of the time compression in a conventional and microwave furnace.
Author's personal copy

304 A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310

Fig. 4. Photographs of the 90W–7Ni–3Fe compacts sintered in (a) conventional and (b) microwave furnace at temperatures ranging between 1200 ◦ C and 1450 ◦ C.

maintained the same, the densification parameter follows the sim- The anisotropy in the shrinkage has been attributed to the asym-
ilar trend as sintered density. For all the sintering temperatures metry of the pore shape and its curvature that results due to the
microwave sintering resulted in enhanced densification. In fact, compaction of the powders. As a result the radial direction under-
microwave sintered compacts attained nearly full (>98.5%) density goes more shrinkage due to pore rounding. From Fig. 6, it can also be
at 1450 ◦ C. The increase in densification with increasing tempera- inferred that the above trend (i.e. more radial shrinkage) is more
ture can be attributed to the greater diffusivity of W in the binder. It accentuated in case of microwave sintered compacts. While the
has been reported [25,26] that the diffusivity of tungsten in nickel exact reason for this is not clear, however, based on the modeling
shows a drastic increase from 10−10 cm2 /s to about 10−5 cm2 /s as results [22] and literature report [29], it is hypothesized that the
the temperature increases from 1200 ◦ C to 1350 ◦ C. curvature difference of the pores in the cylindrical compacts may
Fig. 6a and b show the variation in the radial and axial shrinkage have resulted in differential microwave coupling, which, in turn,
with sintering temperature for the 90W–7Ni–3Fe alloys consol- may result in asymmetric dimensional change during heating.
idated in conventional and microwave furnace, respectively. As To understand phenomenology of densification during sinter-
evident from the figure, for both the heating modes, shrinkage of ing, it is important to first evaluate the phase changes that occur in
the cylindrical compacts in radial direction is invariably higher than the premixed W–Ni–Fe powder compacts during thermal cycling.
that measured axially. The shrinkage trend correlates well with the To isolate the influence of tungsten addition on the associated phase
densification results summarized in Fig. 5a. As compared to the changes, DSC analysis was conducted separately for 70Ni–30Fe
conventional sintering, the dimensional shrinkage in both radial as and 90W–7Ni–3Fe system. The corresponding DSC curves are pre-
well as axial direction is higher for compacts sintered in microwave sented in Figs. 7 and 8, respectively. For the 90W–7Ni–3Fe compact,
furnace. Elsewhere, similar results have been reported in other sys- two sequential endothermic peaks are formed at 1419 ◦ C and
tems as well by Lenel et al. [27] and El-Shanshoury and Nazmy [28]. 1441 ◦ C during heating (Fig. 7a), and the corresponding exother-

Fig. 5. Effect of sintering temperature on the (a) density and (b) densification parameter of 90W–7Ni–3Fe alloys consolidated in conventional and microwave furnace.
Author's personal copy

A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310 305

Fig. 6. Effect of sintering temperature and heating mode on the (a) axial and (b) radial shrinkage of 90W–7Ni–Fe alloys.

mic peak upon cooling occurs at 1436 ◦ C and 1445 ◦ C, respectively perature of the Ni–Fe matrix is well above the melting point of
(Fig. 7b). According to the Fe–Ni phase diagram [30], iron and elemental nickel and does not significantly change with the in situ
nickel are mutually soluble over a wide temperature range. The alloying among elemental Ni, Fe, and W powders. This agrees well
30:70 ( Fe, Ni) alloy melts congruently at 1445 ◦ C. The endother- with the equilibrium conditions in Fe–Ni and Fe–Ni–W phase dia-
mic peak at 1441 ◦ C is related to the melting temperature of grams [30,32]. According to the Fe–Ni phase diagram [30], iron and
Ni–Fe alloys. The depression in the solidification temperature under nickel are mutually soluble over a wide temperature range. The
non-equilibrium processing condition is well known [31]. Com- 30:70 ( Fe, Ni) alloy melts congruently at 1445 ◦ C. In the Fe–Ni–W
pared with 7Ni–3Fe, the 90W–7Ni–3Fe compact exhibits a much ternary alloys phase diagrams [32], ( Fe, Ni) also acts as one of the
similar temperature range for liquid-phase formation and solidifi- main phases between 800 ◦ C and 1500 ◦ C, and the eutectic Ni–Fe–W
cation. During heating (Fig. 8a), the endothermic peak at 1477 ◦ C liquid phase forms at 1465 ◦ C. Besides liquid-phase formation, the
represents the melting of the Ni–Fe matrix. In accordance, the sintering process in the W–Ni–Fe alloy system may also involve
exothermic peak appears at 1471 ◦ C (Fig. 8b) corresponds to the some other phase changes, most of which relate to the forma-
solidification of the binder melt. It is evident that the melting tem-

Fig. 8. DSC analysis of 90–7Ni–3Fe powder mixture during (a) heating and (b) cool-
Fig. 7. DSC analysis of 70Ni–30Fe alloy powder during (a) heating and (b) cooling ing at a constant rate of 10 ◦ C/min. The powders were heated up to 1500 ◦ C in
at a constant rate of 10 ◦ C/min. The powders were heated up to 1500 ◦ C in argon. argon.
Author's personal copy

306 A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310

Fig. 9. Dilatometric shrinkage and shrinkage rate curve of 90W–7Ni–3Fe compacts


heated up to 1500 ◦ C at a constant heating rate of 10 ◦ C/min.

tion of intermetallics [33,34]. As evident in Fig. 8, an endothermic


peak forms at 1253 ◦ C for the 90W–7Ni–3Fe compact during heat-
ing (Fig. 6a), and an exothermic peak appears at 1174 ◦ C upon its
cooling (Fig. 8b). These two peaks are very likely resulted from the
transformation of Fe7 W6 , which is the sole intermetallic phase in
the Fe–Ni–W alloy system above 1150 ◦ C [32]. The results from
the present study correspond well with those reported elsewhere
[12,35,36]. Unlike the previous reports [33,34], the alloy used in the
present study had a stoichiometric binder composition (Ni:Fe 7:3),
hence, formation of the intermetallic is not expected. However, as
the composition in the present study was prepared by powder mix-
ing, it is likely that this may have caused some local inhomogeneity Fig. 10. Dilatometric plots showing the effect of heating rate on the (a) linear shrink-
that may have resulted in formation of small amounts of inter- age and (b) shrinkage rate of 90W–7Ni–3Fe alloy heated up to 1500 ◦ C in reducing
atmosphere.
metallic phases which sensitive equipment such as DSC can readily
detect. To ensure good homogeneity, it would be interesting to con-
duct further studies using prealloyed Ni–Fe powders instead of the ature measurement and the differences in the atmosphere used,
premixed ones. this closely corresponds with the melt formation temperature of
From the DSC curves, it can be inferred that out of the five tem- this alloy determined using DSC (Fig. 8) and reported by others
peratures selected for conventional sintering, lower temperatures [12,35,36]. The enhanced densification at the onset of melting is
(1200 ◦ C, 1250 ◦ C and 1300 ◦ C) correspond to solid-state sintering, due to the capillary-induced arrangement. In addition, melt forma-
and the higher ones (1450 ◦ C and 1500 ◦ C) to liquid phase sintering. tion also results in enhanced tungsten solubility. The tungsten is
It is well known that due to faster diffusion rate and capillary-
induced stresses the densification kinetics is higher during liquid
phase sintered compacts [37]. The densification response of the
alloys in the present study (Fig. 5b) confirms that they are in accor-
dance with this. Another factor that contributes to densification
during liquid phase sintering is that the newly formed wetting melt
spreads to fill the small pores and preferentially penetrates the
grain boundaries [37,38]. Since tungsten has substantial solubility
into the Ni–Fe matrix, hence the melt tends to dissolve the interpar-
ticle necks and spreads out [39], this results in the decrease of the
solid–liquid interfacial energy below that of the equilibrium value
[40] and further contributes to densification in the initial period
just after melt formation.
To obtain a better insight about the densification response of
the 90W–7Ni–3Fe compacts, Fig. 9 plots the shrinkage and shrink-
age rate data of the alloy at various temperatures during heating at
a rate of 10 ◦ C/min in a dilatometer. The compact did not show
any significant dimensional change till about 1000 ◦ C. From the
figure, it is quite evident that up to 900 ◦ C the compact does not
undergo any significant shrinkage. Subsequently, a substantial por-
tion of the densification (10% linear shrinkage) occurs prior to the
melt formation. This is in agreement with the similar observations
Fig. 11. The peak shrinkage rate of 90W–7Ni–3Fe alloy at the onset of liquid forma-
reported by other researchers [41–43]. From the shrinkage rate
tion temperature plotted as a function of heating rate. The temperature at which the
curve (Fig. 9) it is evident that there is a sudden increase in the peak shrinkage occurs is also indicated on the plot. The data from Bollina and German
densification at 1463 ◦ C. Taking into account the error in temper- [12] on 88W–Ni–Fe (7:3) have also been superimposed for the sake of comparison.
Author's personal copy

A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310 307

Fig. 12. SEM micrographs of 90W–7Ni–3Fe compacts sintered in a conventional furnace (left) and microwave furnace (right) at (a) 1200 ◦ C, (b) 1250 ◦ C, 1300 ◦ C and (b)
1450 ◦ C, respectively.

preferentially dissolved from the interparticle necks and/or sharp In the above section, dilatometric investigation aided in under-
curvatures. This results in disruption of the solid-state sintered standing the influence of temperature on the phenomenology of
W–W bonds and causes grain shape accommodation. Both results densification. Besides temperature, heating rate is also an impor-
in activating densification kinetics and accounts for the greater tant parameter that influences the densification since it influences
shrinkage rates observed in Fig. 9. not only the diffusivity but also alters the phase transforma-
Author's personal copy

308 A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310

tion temperature. Unfortunately, owing to the limitation of the


dilatometer used in the present investigation, the fast heating
rates obtained in microwave heating could not be achieved. Hence,
to evaluate the influence of heating rate on the sinterability, the
90W–7Ni–3Fe compacts were heated up to 1500 ◦ C at three heat-
ing rates, namely 2 ◦ C/min, 5 ◦ C/min and 10 ◦ C/min in a dilatometer.
The corresponding shrinkage and shrinkage rate variation at the
three heating rates are summarized in Fig. 10a and b, respectively.
It is interesting to note that the contribution to densification from
solid-state sintering is higher at lower heating rates (Fig. 10a). In the
temperature range corresponding to onset of melting, the effect of
heating rate is better reflected in the shrinkage rate plot (Fig. 10b)
since it is more sensitive to small changes in temperature. Inter-
estingly, the contribution of liquid phase sintering to densification
is higher for alloys sintered at higher heating rate. Similar effect
of heating rate on the shrinkage and shrinkage rate were observed
by Bollina and German [12] in 88W–Ni–Fe alloy (Ni/Fe ratio 7:3). Fig. 13. Representative microstructure of 90W–7Ni–3Fe alloy liquid phase sintered
at 1500 ◦ C in a conventional furnace.
The shrinkage rate results near binder melting temperature (plot-
ted as inset to Fig. 10b) reveals that both the shrinkage rate and
matrix melting temperature shift to higher values as the heating expected and correlate well with the DSC (Fig. 7) and dilatometric
rate increases [Fig. 10b (inset)]. Fig. 11 shows the peak tempera- (Fig. 8) observations. However, the corresponding microstructures
ture corresponding to the maximum shrinkage rate in the present of 90W–7Ni–3Fe alloys microwave sintered at 1300 ◦ C and 1450 ◦ C
alloy at various heating rates and along with the results from Bollina (Fig. 12c and d) are very intriguing. The samples microwave sin-
and German [12] in a similar alloy system. tered at these temperatures show significantly coarse tungsten
At any temperature the shrinkage rate is lower for a slower grains and the microstructure is similar to that expected of a typ-
heating rate. The point of maximum shrinkage rate indicates the ically liquid phase sintered W–Ni–Fe alloy! In view of the fact
dissolution of solid-solid bonds by liquid phase and densification that the matrix melting temperature in these alloys is elevated as
due to rearrangement. If the solid-state densification is suppressed, the heating rate increases (Fig. 11), this is even more perplexing.
then melt formation significantly contributes to densification of This finding was validated by several repeated experiments. These
the compact. The shrinkage at any given temperature is higher for results are similar to those reported in 90W–7Ni–3Cu system [19]
the slower heating rate of 2 ◦ C/ min, since the slower heating rate and need to be more critically investigated.
gives more time for solid-state diffusion prior to liquid formation. One of the probable cause of this observation can be attributed
As the compact becomes more dense prior to reaching the liquid to the inherent problem in accurately measuring temperature
formation temperature, hence the contribution from liquid phase during microwave sintering. The temperature measurement for
sintering proportionately decreases at slower heating rates [44]. By the present study was done by optical pyrometer. To address
extending this argument it is hypothesized that during microwave this inherent limitation, temperature was also measured using a
sintering the major contribution to densification stems from liquid more accurate emmissivity-based infra-red pyrometer, wherein,
phase sintering. It is envisaged that the fast heating rates achieved the emmissivity at various elevated temperatures was calibrated
through microwaves restrict the W–W bond formation prior to (for the same composition compacts) in a conventional furnace.
melting and thereby making the contribution to densification from Even in this, same results were obtained. One of the probable rea-
rearrangement more effective. Elsewhere too, researchers [44–46] sons for this observation is also the critique that the pyrometer
have proposed that slow heating rates are detrimental to rearrange- based temperature measurements are an average of the surface
ment. temperature only. It may be likely that – for multi-component sys-
tem, such as W–Ni–Fe alloys – the temperature in the interior of
3.3. Microstructural evolution during conventional and the sample may be slightly higher. However, no gradient in the
microwave sintering microstructure was obtained in the sintered samples. One of the
likely explanations for such an observation therefore could be the
Fig. 12a–d compare the effect of heating mode on the differential coupling of the constituents with microwave thereby
microstructure of the 90W–7Ni–3Fe compacts sintered at 1200 ◦ C, leading to anisothermal heating effect that has been reported to
1250 ◦ C, 1300 ◦ C, and 1450 ◦ C, respectively. At 1500 ◦ C, the cause rapid diffusion rates and onset of liquid phase sintering at
W–Ni–Fe compacts were only conventionally sintered and the cor- lower temperatures [47,48]. Besides microstructural coarsening,
responding microstructure is presented in Fig. 13. As expected, as a consequence of this enhanced diffusion, the densification is
for both conventional as well microwave heating, higher sinter- expected to be high, which is indeed the case (Fig. 5). This is further
ing temperatures, result in reduced porosity, more homogenous confirmed by the EPMA results summarizing the effect of heating
distribution of the binder phase and coarser tungsten grains. mode and sintering temperature on the composition of the matrix
For compacts sintered at 1200 ◦ C and 1250 ◦ C, the microstruc- phase. Note that irrespective of the temperature, the amount of dis-
tures are typical of those expected from solid-state sintering. In solved tungsten in microwave sintered alloy is higher than those
case of conventional sintering, the compacts sintered at 1300 ◦ C, consolidated conventionally (Table 2).
again corresponds to solid-state sintering. In comparison, while Fig. 14a and b compare the XRD patterns of the samples sintered
the microstructure of compacts conventionally sintered at 1450 ◦ C at various temperatures in conventional and microwave furnace,
exhibit significant grain coarsening, the tungsten grains still have respectively. As discussed earlier, the choice of a stoichiometric
irregular shape which underscores solid-state sintering. Further binder ratio (7:3) obviates intermetallic formation in any of the
increase in the sintering temperature to 1500 ◦ C results in a typ- sintered compact. The microwave sintered 90W–7Ni–3Fe alloys are
ical liquid phase sintered microstructure with well-rounded and characterized by a change in the peak intensity and absence of some
less contiguous tungsten grains interspersed in a Ni–Fe matrix of the matrix diffraction peaks. Such observations have also been
(Fig. 13). These results when looked in isolation are as per noticed in other systems as well and have been attributed to the
Author's personal copy

A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310 309

Table 2
Effect of heating mode and sintering temperature on the composition of the matrix phase in 90W–7Ni–3Fe alloys.

Temperature, ◦ C Heating mode Elemental content, wt. %

W Ni Fe

Conventional 8.11 ± 5.85 24.65 ± 17.85 67.48 ± 24.34


1200 ◦ C
Microwave 19.14 ± 10.9 45.24 ± 12.41 35.63 ± 24.62

Conventional 27.38 ± 5.8 57.67 ± 0.81 11.21 ± 6.35


1250 ◦ C
Microwave 30.23 ± 4.58 58.45 ± 0.38 10.18 ± 5.17

Conventional 21.04 ± 3.65 53.96 ± 4.00 23.19 ± 7.48


1300 ◦ C
Microwave 28.43 ± 0.5 51.07 ± 0.12 20.65 ± 0.41

Conventional 23.19 ± 1.05 52.63 ± 0.58 21.3 ± 0.34


1450 ◦ C
Microwave 25.62 ± 0.66 51.48 ± 0.28 21.7 ± 0.27

Conventional 22.96 ± 1.71 52.5 ± 0.76 21.77 ± 0.34


1500 ◦ C
Microwave – – –

Table 3
Effect of heating mode and sintering temperature on the average grain size, transverse rupture strength (TRS) and hardness (bulk and micro) of the 90W–7Ni–3Fe compacts.

Sintering temperature , ◦ C Heating mode Grain size, ␮m Hardness HV5 Vickers microhardness TRS MPa

WHV0.05 MatrixHV0.015

CON – 161 ± 27 – – 212


1300
MW 17 ± 7 325 ± 15 – – 751

CON 18 ± 11 289 ± 23 – – 782


1450
MW 20 ± 7 352 ± 11 475 ± 15 403 ± 14 1800 ± 11

1500 CON 26 ± 14 329 ± 17 454 ± 8 402 ± 14 1796 ± 19

creation of defects upon interaction with the microwaves [48,49].


The presence of such defects and their nature needs to be investi-
gated further through transmission electron microscopy (TEM) and
orientation imaging microscopy (OIM/EBSD) techniques.

3.4. Mechanical properties of sintered W–Ni–Fe alloys

Table 3 summarizes the effect of heating mode and sintering


temperature on the average grain size, bulk and micro-hardness
and the flexural strength of the 90W–7Ni–3Fe alloys. Owing to
the fine and contiguous microstructure, it is rather difficult to
accurately measure the tungsten grain size in solid-state sintered
compacts. For conventional sintering, the grain size measurements
were performed for samples sintered at 1450 ◦ C and 1500 ◦ C. In
case of microwave sintering, the grain sizes were quantified for
compacts sintered at 1300 ◦ C and 1450 ◦ C. Note that as compared to
conventionally liquid phase sintered alloys, the grain size was lower
for compacts consolidated in microwave furnace. This, along with
a higher sintered density, resulted in significantly higher hardness
and the bend strength in microwave sintered alloys. The hardness
values are similar or better than those reported previously [50,51].
The higher transverse rupture strength in microwave sintered sam-
ples can be attributed with the spheroidal tungsten grains which
are less contiguous and hence provide higher mean free path for
crack propagation [52–54]. It is well-recognized that the mechan-
ical properties of the as-sintered W–Ni–Fe alloys can be further
improved by heat-treatment [55,56]. In view of the high tungsten
solubility in the matrix, anomalous microstructural and phase evo-
lution during sintering, a future extension of the present set of study
could be to evaluate critically the detailed DSC analysis of the com-
position and also the effect of thermo-mechanical treatment on
the mechanical properties of microwave sintered tungsten heavy
alloys.

4. Conclusions
Fig. 14. XRD plots showing the effect of sintering temperature on the phase evolu-
tion in 90W–7Ni–3Fe alloy sintered in (a) conventional and (b) microwave furnace. In this study, 90W–7Ni–3Fe alloys were successfully sin-
tered both through solid-state as well as liquid phase sintering
Author's personal copy

310 A. Mondal et al. / Journal of Alloys and Compounds 509 (2011) 301–310

in microwave furnace. As compared to conventional sintering, [17] A. Mondal, A. Upadhyaya, D. Agrawal, J. Microwave Power Electromagn. Energy
microwave sintering resulted in about 80% reduction in the overall JMPEE 43 (2009) 11–16.
[18] S.D. Luo, J.H. Yi, Y.L. Guo, Y.D. Peng, L.Y. Li, J.M. Ran, J. Alloys Compd. 473 (2009)
processing time. To evaluate the influence of heating rate on the sin- L5–L9.
terability, the 90W–7Ni–3Fe compacts were heated up to 1500 ◦ C [19] A. Mondal, A. Upadhyaya, D. Agrawal, Mater. Sci. Eng. A 527 (2010) 6870–6878.
at three heating rates, namely 2 ◦ C/min, 5 ◦ C/min and 10 ◦ C/min in [20] C. Zhou, J. Yi, S. Luo, Y. Peng, L. Li, G. Chen, J. Alloys Compd. 482 (2009) L6–L8.
[21] A. Mondal, D. Agrawal, A. Upadhyaya, J. Microwave Power Electromag. Energy
a dilatometer. The W–Ni–Fe alloys exhibit significant densification JMPEE, 44, in press.
prior to melt formation through solid-state sintering. The in situ [22] A. Mondal, Modeling of Microwave Heating of Particulate Metals and its Appli-
dilatometric studies reveal that the contribution to densification cation in Sintering of Tungsten-Based Alloys, Ph.D. Thesis, The Indian Institute
of Technology, Kanpur, 2010.
from solid-state sintering is higher at lower heating rates. By corol- [23] D. Agrawal, J. Cheng, M. Jain, G. Skandan, R. Dowding, K. Cho, B. Klotz, D. Kapoor,
lary, the contribution of liquid phase sintering to densification is Eighth Int. Conf. Sci. Hard Mater., 2004, pp. 143–144.
higher for alloys sintered at higher heating rate. It is also interesting [24] MPIF Standard 41, Standard Test Methods for Metal Powders and Powder Met-
allurgy Products, MPIF, Princeton, New Jersey, USA, 2002, pp. 55–57.
to note that microwave sintered compacts had a well-developed
[25] G.E. Murch, C.M. Bruff, Diffus. Solid Met. Alloys 26 (1990) 279–371.
liquid phase sintered structure at about 50 ◦ C lower temperature [26] J.P. Leonard, T.J. Renk, M.O. Thompson, M.J. Aziz, Metall. Mater. Trans. A 35
as compared to conventional sintering. Accordingly, microwave (2004) 2803–2807.
sintered compacts had about 23% lower tungsten grain size and [27] F.V. Lenel, H.H. Hausner, E. Hayashi, G.S. Ansell, Powder Metall. 8 (1961) 25–36.
[28] I.A. El-Shashoury, M.Y. Nazmy, Powder Metall. 11 (1968) 63–72.
22% improvement in the bulk hardness (352 VHN). As compared to [29] M. Wllert-Porada, Microwave: Theory and Applications in Materials Processing
90W–7Ni–3Fe alloys sintered at 1450 ◦ C in a conventional furnace, IV, 1998, pp. 158–164.
microwave sintering led to significant (∼130%) improvement in the [30] J.I. Goldstein, Alloy Phase Diagram, ASM Metals Reference Book, American Soci-
ety for Metals, Metals Park, OH, 1981.
transverse rupture strength from 782 MPa to 1800 MPa. [31] D.M. Stefanescu, Science and Engineering of Casting Solidification, 2nd ed.,
Springer, New York, NY, USA, 2009, pp. 1–20.
Acknowledgement [32] P. Villars, A. Prince, H. Okamoto, Handbook of Ternary Alloy Phase Diagrams,
ASM International, Materials Park, OH, 1997.
[33] B.C. Muddle, Metall. Trans. A 15 (1984) 1089–1098.
This collaborative research was done as a part of the Center [34] J.B. Posthill, D.V. Edmonds, Metall. Trans. A 17 (1986) 1921–1934.
for Development of Metal-Ceramic Composites through Microwave [35] S.S. Mani, R.M. German, Advances in Powder Metallurgy, vol. 1, Metal Powder
Industries Federation, Princeton, NJ, 1990, pp. 453–467.
Processing which was funded by the Indo-US Science and Technol- [36] Y. Wu, R.M. German, B. Marx, P. Suri, R. Bollina, J. Mater. Sci. 38 (2003)
ogy Forum (IUSSTF), New Delhi. 2271–2281.
[37] V.L. Yupko, R.V. Minakova, O.P. Kolchin, L.S. Vodopyanova, N.I. Monastyreva,
V.L. Voitenko, Sov. Powder Metall. Met. Ceram. 22 (1983) 44–47.
References [38] R.M. German, P. Suri, S.J. Park, J. Mater. Sci. 44 (2009) 1–39.
[39] L. Ekbom, A. Eliasson, Processing of Powder Metallurgy World Congress PM’94,
[1] A. Upadhyaya, Trans. Indian Inst. Met. 55 (2002) 51–69. vol. II, European Powder Metallurgy Association, Shrewsbury, UK, 1994, pp.
[2] E. Lassner, W.D. Schubert, Tungst: Properties, Chemistry, Technology of the 1565–1572.
Elements, Alloys, Chemical Compounds, Kluwer Academic/Plenum Publishers, [40] I.A. Aksay, C.I. Hoge, J.A. Pask, J. Phys. Chem. 78 (1974) 1178–1183.
New York, NY, USA, 1999. [41] A. Upadhyaya, J. Mater. Chem. Phys. 67 (2001) 101–110.
[3] V. Srikanth, G.S. Upadhyaya, Int. J. Refract. Hard Met. 5 (1986) 49–54. [42] W.E. Gurwell, in: A. Bose, R.J. Dowding (Eds.), Tungsten and Refractory Metals,
[4] R.M. German, Liquid Phase Sintering, Plenum Press, NY, 1985. vol. 2, Metal Powder Industries Federation, Princeton, NJ, 1994, pp. 65–75.
[5] I.Y. Dzykovich, R.V. Makarova, O.K. Teodorovich, I.N. Frantsevich, Sov. Powder [43] F. Akhtar, Int. J. Ref. Met. Hard Mater. 26 (2008) 145–151.
Metall. Met. Ceram. 4 (1965) 655–660. [44] R.M. German, S. Farooq, in: S. Sōmiya, M. Shimada, M. Yoshimura, R. Watanabe
[6] S.G. Caldwell, in: A. Bose, R.J. Dowding (Eds.), Tungsten and Tungsten (Eds.), Sintering’87, vol. 1, Elsevier Science Publishing Co. Inc., New York, NY,
Alloys—1992, Metal Powder Industries Federation, Princeton, NJ, 1992, pp. 1988, pp. 459–464.
89–96. [45] G.J. Shu, K.S. Hwang, Mater. Trans. 46 (2005) 298–302.
[7] B.H. Rabin, A. Bose, R.M. German, Int. J. Powder Metall. 25 (1989) 21–27. [46] B.D. Storozh, P.S. Kislyi, Powder Metall. Met. Ceram. 13 (2005) 712–716.
[8] A. Belhadjhamida, R.M. German, in: E.S. Crowson, Chen (Eds.), Tungsten and [47] R.D. Peelamedu, R. Roy, D. Agrawal, Mater. Res. Bull. 36 (2001) 2723–2739.
Tungsten Alloys—Recent Advances, The Minerals, Metals & Materials Society, [48] R. Roy, Y. Fang, J. Cheng, D.K. Agrawal, J. Ceram. Am. Soc. 88 (2005) 1640–1642.
Warrendale, PA, USA, 1991, pp. 3–19. [49] R. Peelamedu, R. Roy, L. Hurtt, D. Agrawal, A.W. Fliflet, D. Lewis III, R.W. Bruce,
[9] L.L. Bourguignon, R.M. German, Int. J. Powder Metall. 24 (1988) 115–121. Mater. Chem. Phys. 88 (2004) 119–129.
[10] A. Bose, D. Sims, R.M. German, Metall. Trans. A 19 (1988) 487–494. [50] V. Srikanth, G.S. Upadhyaya, J. Mater. Sci. Lett. 7 (1988) 195–197.
[11] R.M. German, A. Bose, S.S. Mani, Metall. Trans. A 23 (1992) 211–219. [51] M. Debata, A. Upadhyaya, J. Mater. Sci. 39 (2004) 2539–2541.
[12] R. Bollina, R.M. German, Int. J. Refract. Met. Hard Mater. 22 (2004) 117– [52] G.S. Upadhyaya, V. Srikanth, Modern Developments in Powder Metallurgy, vol.
127. 17, Metal Powder Industries Federation, Princeton, NJ, 1985, pp. 51–75.
[13] A. Upadhyaya, S.K. Tiwari, P. Mishra, Scripta Mater. 56 (2007) 5–8. [53] I.S. Humail, F. Akhtar, S.J. Askari, M. Tufail, X. Qu, Int. J. Refract. Met. Hard Mater.
[14] M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding, D. Kapoor, D. Agar- 25 (2007) 380–385.
wal, J. Chang, Int. J. Powder Metall. 42 (2006) 53–57. [54] S.H. Islam, X. Qu, S.J. Askari, M. Tufail, X. He, Rare Met. 26 (2007) 200–204.
[15] M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding, D. Kapoor, D. Agar- [55] S.G. Caldwell, Int. J. Powder Metall. 39 (2003) 43–51.
wal, J. Chang, Int. J. Powder Metall. 42 (2006) 45–50. [56] W.H. Baek, M.H. Hong, E.P. Kim, J.W. Noh, S. Lee, H.S. Song, S.H. Lee, Solid State
[16] G. Prabhu, A. Charkraborty, B. Sarma, Int. J. Refract. Met. Hard Mater. 27 (2009) Phenom. 118 (2006) 35–40.
545–548.
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 Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Int. Journal of Refractory Metals and Hard Materials 28 (2010) 597–600

Contents lists available at ScienceDirect

Int. Journal of Refractory Metals and Hard Materials


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / I J R M H M

Effect of heating mode on sintering of tungsten


Avijit Mondal a, Anish Upadhyaya a,⁎, Dinesh Agrawal b
a
Department of Materials and Metallurgical Engineering, Indian Institute of Technology, Kanpur 208016, India
b
The Pennsylvania State University, University Park, PA 16802, USA

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

Article history: Microwave heating is recognized for its various advantages, such as time and energy saving, very rapid
Received 4 February 2010 heating rates, considerably reduced processing cycle time and temperature, fine microstructures and
Accepted 5 May 2010 improved properties. The present paper investigates the feasibility of consolidating tungsten powders
through microwave sintering. A comparative analysis has also been attempted between the sintering
Keywords: response of pure tungsten powder compact in a microwave and conventional furnace.
Microwave sintering
© 2010 Elsevier Ltd. All rights reserved.
Tungsten
Microstructures

1. Introduction enhanced solubility and mass transport kinetics. The additives which
remain segregated to the W–W powder interface and in which tungsten
Tungsten belongs to Group VIB of periodic table. As a refractory has appreciable solubility, act as short-circuit diffusion pathways,
metal, tungsten is characterized by its very high melting point thereby promoting densification.
(3420 °C), high density (19.3 g/cm3), low coefficient of thermal Another important approach to activate sintering of tungsten is
expansion (4.4 ppm/K at 20 °C) and superior mechanical properties through selection of submicron or nano-sized precursor tungsten
at elevated temperatures, which render it highly suitable for many powder. However, such powders are expensive and are prone to
engineering applications such as lighting filaments, heating source, contamination [7]. Many studies have shown that sintering temper-
aerospace, electronic devices, sports and military uses, etc. [1]. Owing ature is related to the powder size, when the size is in nano-scale, the
to very high fusion point, the consolidation of a conventional sintering temperature can be decreased up to several hundreds of
microcrystalline W powder is difficult and generally requires a degrees. The reduction of sintering temperature for nano tungsten has
temperature in excess of 1700 °C through solid-state sintering in been reported by several researchers [7–10]. Bose et al. [7] have also
electrical resistance sintering furnace under hydrogen atmosphere. shown that pressure assisted process such as plasma pressure
Achievement of near theoretical sintered densities for pure tungsten compaction helps in the reduction of process temperature. The
at temperatures below 1650 °C is typically not feasible [2]. Densification reported sintering temperature of nano-sized tungsten produced by
of refractory metals, such as tungsten, can be enhanced greatly by high energy mechanical milling was drastically decreased from
activating the sintering process wherein the sintering temperature is conventional temperature of 2500 °C to 1700 °C [8]. Other processes
appreciably lowered. Activated sintering refers to combination of such as hot-isostatic processing [10], and spark plasma sintering [11],
processing approaches that reduce the activation energy for sintering. too result in further reduction in the processing temperature.
One such technique to activate sintering is by addition of small amounts Because of the characteristic feature and obvious advantages,
(b1 wt.%) of Group VIII transition metals [3–6]. It has been reported that application of microwave energy in consolidating particulate materi-
the sintering temperature of tungsten can be brought down from als has become a preferred method over conventional (resistant
2800 °C to 1400 °C by less than 1 wt.% addition of transition metals, such heating) technique. The application of microwave energy in consol-
as palladium and nickel [6]. German and Munir [5,6] have extensively idating tungsten powders was first studied by Jain et al. [12]. A
investigated the role of various transition metal additions in densifica- comprehensive study on microwave sintering of tungsten and its
tion activation of tungsten powder compacts. From their study, it is quite alloys were also conducted by several researchers [13–21]. This paper
evident that some transition metals (e.g. Pd, Ni) enhance densification, reports the consolidation of nano-sized tungsten powder through
while others (e.g. Ag, Cr) have little influence. Densification enhance- microwave and conventional sintering methods. To evaluate the
ment in tungsten by transition metal additives was attributed to interaction of microwaves, tungsten powder with a median particle
size of 72 nm was subjected to 2.45 GHz microwaves in a multimode
furnace. For comparing the effect of heating mode, in a parallel set of
⁎ Corresponding author. Tel.: + 91 512 2597672; fax: + 2597505. experiments, tungsten powder compacts pressed to similar green
E-mail address: anishu@iitk.ac.in (A. Upadhyaya). density levels were consolidated in a conventional furnace.

0263-4368/$ – see front matter © 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijrmhm.2010.05.002
Author's personal copy

598 A. Mondal et al. / Int. Journal of Refractory Metals and Hard Materials 28 (2010) 597–600

2. Experimental procedure quite rapidly. Tungsten powders owing to their fine size are
susceptible to oxidation and hence need to be processed in reducing
Tungsten of average crystallite size of 72 nm powders was (hydrogen) atmosphere. Due to the poor thermal shock resistance of
supplied by NEI Corporation, New Jersey, USA. The as-received the alumina tube used, the heating rate of the conventional sintering
powders were pressed in a die of 1.6 cm inner diameter to make the was restricted to 5 °C/min. Furthermore, to ensure homogenization of
green compacts of approximately 0.3 to 0.4 cm in height. The green the temperature, isothermal hold at intermediate sintering tempera-
density was around 49% of theoretical. To study the densification tures was provided. Unlike conventional sintering, in a microwave
behavior the green (as-pressed) compacts were sintered using furnace, the tungsten powder compact per se acts as a source of heat
conventional and microwave furnace. The conventional sintering of since it couples directly with microwaves. Consequently, the overall
green compacts was conducted in a MoSi2 heated horizontal tubular heating rate achieved in microwave furnace was ∼25 °C/min for the
sintering furnace (model: OKAY 70T-7, supplier: Bysakh, Kolkata, tungsten compacts. Taking into consideration the lower heating rate
India). Microwave sintering of the green compacts was carried out and the intermittent isothermal holds in conventional furnace
using a multimode cavity 2.45 GHz, 6 kW microwave furnace. Further sintering, there is about 90% reduction in the overall processing
details of the microwave furnace and experimental arrangements time during microwave sintering.
have been described elsewhere [19]. For each set of experiments Unlike ceramic materials microwave interaction with metals is
(conventional and microwave sintering) four samples were investi- restricted to its surface only. This depth of penetration in metals, also
gated and the as-sintered samples were characterized for sintered known as skin depth (δ), is defined as the distance into the material at
density through both dimensional measurements as well as Archi- which the incident power drops to 1/e (36.8%) of the surface value.
medes' density measurement techniques. To take into account the The skin depth is mathematically expressed as follows:
influence of the initial as-pressed density, the compact sinterability
was also expressed in terms of densification parameter which is 1
δ = pffiffiffiffiffiffiffiffiffiffiffiffi ð2Þ
calculated as follows: πf μσ

ðsintered density−green densityÞ where, f is the microwave frequency (2.45 GHz), μ is the magnetic
Densification parameter = ð1Þ
ðtheoretical density−green densityÞ permeability, σ is the electrical conductivity, and ρ is the electrical
resistivity of the metals. From Eq. (2), it is evident that metals with
Metallographic techniques were employed on the sintered higher electrical conductivity have lower skin depths. For metals, as
samples. The sintered samples were wet polished in a manual the resistivity increases with increase in temperature, the skin depth
polisher (model: Lunn Major, supplier: Struers, Denmark) using a too increases. Resistivity as a function of temperature has been
series of 6 µm, 3 µm and 1 µm diamond paste, followed by cloth considered from the literature [22] and Fig. 2 plots the effect of
polishing using a suspension of 0.04 µm colloidal SiO2 suspension. temperature on the skin depth of tungsten. For the sake of comparison
Murakami's reagent was used for etching the sintered tungsten similar graphs have been plotted for various other conductive metals
samples. The scanning electron micrographs of as-polished samples as well. It is interesting to note as compared to some other metals, the
were obtained by a scanning electron microscope (model: FEI quanta, skin depth of tungsten is relatively higher. Since the initial tungsten
Netherlands) in the secondary electron (SE) and back scattered (BSE) powder size is lower hence the compacts undergo relatively
mode. Bulk hardness measurements were performed on polished homogenous heating despite the heating rate being so high. Since
surfaces of sintered cylindrical compacts at a load of 5 kg using Vickers the objective of the present work was to restrict the temperature to
hardness tester (model: V100-C1, supplier: Leco, Japan). 1600 °C, owing to the small size of the initial powder the temperature
was achieved in just applying 900 W of power. In comparison, to heat
3. Results and discussion the similar compact in a conventional furnace wherein the heat
transfer was indirect and limited by the radiative losses, the input
3.1. Effect of heating mode on sintering of tungsten power was on an average 2.1 kW. While it is possible to increase the
heating rate in the conventional sintering by increasing the power
Fig. 1 compares the thermal profiles for tungsten powder compacts throughput, it will be at the expense of overall furnace life. One such
heated in conventional and microwave furnace. It is interesting to experiment was tried; however, the as-pressed compact could not
note that W compact couples with microwaves and gets heated up withstand the thermal gradient and pulverized during processing.
Recently, Prabhu et al. [14] too have reported microwave sintering
of tungsten powders prepared by high energy milling. However, as
compared to our study, they could achieve 93% theoretical density at
1800 °C with a heating rate of 7.8 °C/min and at a power throughput

Fig. 2. The variation in the skin depth of tungsten subjected to 2.45 GHz microwave. For
Fig. 1. Comparison of the thermal profiles of tungsten compact sintered in a radiatively- comparison, the skin depths of some common metals have been plotted for various
heated (conventional) and microwave furnace. temperatures.
Author's personal copy

A. Mondal et al. / Int. Journal of Refractory Metals and Hard Materials 28 (2010) 597–600 599

Fig. 3. Effect of heating mode on (a) the sintered density and densification parameter and (b) the axial and radial shrinkage of tungsten compact. All compacts were pressed at
100 MPa and sintered at 1600 °C in reducing atmosphere.

of 16 kW. Furthermore, they did not compare their sintering results particle sizes the onset of densification occurs at lower temperatures
with the conventionally sintered W powder compacts. of 0.2–0.4Tm (Tm — melting temperature in K) as compared to 0.5–
0.8Tm for micron-sized powders [23]. Moreover, it has been
3.2. Densification response of sintered tungsten established that for nano-structured powders, grain boundary sliding
and rotation and viscous flow significantly contribute to densification
Fig. 3a compares the effect of heating mode on the sintered density enhancement [24–27]. In the present study, due to the chemical
and densification parameter of tungsten powder compact. It is worth processing route, the tungsten powders were prepared in nano-scale
noticing that irrespective of the heating mode the compacts undergo in strain free condition. The sintering temperature selected for the
significant densification during sintering. This can be attributed to the present study corresponds to 0.46Tm.
ultra-fine particle size of the powder used for this investigation. Jain et al. [12] also reported about 10 to 12% higher density in case
Elsewhere, Wang et al. [10] too reported similar sintering response of of microwave sintered compacts over the conventional heating. More
ultra-fine sized tungsten powder. The sintered density of the W recently, Prabhu et al. [14] have also investigated microwave sintering
powder compacts in microwave increased from 90% to 95% for pure of high energy milled tungsten powder compacts. However, in their
tungsten compacts when compared with conventional heating. The experiments the compacts were sintered under nitrogen atmosphere
densification parameter is a normalized parameter which takes into and at a high temperature (1800 °C). Despite consolidating at such a
consideration the initial variation of the green density on the high temperature, the sintered density was less than that achieved in
densification response. Since, the initial density of the W powder this study. It can therefore be inferred that for effective utilization of
compact was kept the same hence the densification parameter follows the microwave heating, the powder compacts must be sintered in
a similar trend as the sintered density. reducing atmosphere.
Usually, the sintering characteristics of the tungsten powders have Fig. 3b shows the axial and radial shrinkage of the cylindrical
been investigated in micron-sized powders [10]. For achieving small tungsten compacts sintered in a conventional and microwave furnace.
sizes, these powders were usually milled [8]. Due to the contamina- Due to the low green density (49%), the compacts heated in both the
tion from the milling media and high defect structure within the modes undergo substantial shrinkage. In case of conventional
crystallites, proper evaluation of the sintering response of the sintering, however, the radial shrinkage is more than that observed
tungsten powder is usually difficult. However, the general consensus in the longitudinal (axial) direction. In contrast, the shrinkage in both
in the literature has converged to the fact that for submicron/nano the directions is nearly similar for microwave sintered compacts

Fig. 4. SEM photomicrograph of (a) the as-pressed tungsten powder and the powder compact sintered at 1600 °C for 30 min in (b) conventional and (c) microwave furnace.
Author's personal copy

600 A. Mondal et al. / Int. Journal of Refractory Metals and Hard Materials 28 (2010) 597–600

Table 1 Acknowledgement
Average grain size and bulk hardness of conventionally and microwave sintered
tungsten powder compacts.
The authors would like to acknowledge NEI Corporation, New
Heating mode Grain size, µm Hardness, HV5 Jersey, USA for supplying the powder used for this investigation. This
Conventional 2.9 ± 1.1 412 ± 12 study was also partially supported by the Indo-US Science and
Microwave 2.6 ± 1.0 441 ± 9 Technology Forum (IUSSTF), New Delhi and the Institute for Plasma
Research (IPR), Ahmedabad.

which indicates more isotropic shrinkage behavior. This further References


confirms that the compact heating in microwaves is more uniform
[1] Lassner E, Schubert WD. Tungsten: properties, chemistry, technology of the
despite the high heating rate. This is attributed to the volumetric elements, alloys, and chemical compounds. New York, NY, USA: Kluwer Academic/
heating microwave provides. Plenum Publishers; 1999.
[2] German RM. Critical developments in tungsten heavy alloys. In: Bose A, Dowding
RJ, editors. Tungsten and tungsten alloys — 1992. Princeton, NJ, USA: Metal
3.3. Effect of heating mode on microstructure and hardness Powder Industries Federation; 1992. p. 3–13.
[3] Hayden HW, Brophy JH. The activated sintering of tungsten with Group VIII
elements. J Electrochem Soc 1963;110(7):805–10.
To investigate the effect of heating mode on the microstructural [4] Shaler AJ. Activated sintering — a review. In: Kuczynski GC, editor. Sintering and
coarsening and its homogeneity, the as-sintered compacts were related phenomena. New York, NY, USA: Gordon and Breach; 1967. p. 807–23.
metallographically prepared. Fig. 4a shows the SEM micrograph of the [5] German RM, Munir ZA. Systematic trends in the chemically activated sintering of
tungsten. High Temp Sci 1976;8:267–80.
as-pressed tungsten powder. Fig. 4b and c shows the photomicro- [6] German RM, Munir ZA. Enhanced low-temperature sintering of tungsten. Metall
graphs of compacts sintered in conventional and microwave furnace, Mater Trans A 1976;7A:1873–7.
respectively. It is worth noticing that as owing to the initial low as- [7] Bose A, Klotz BR, Kellogg FR, Cho KC, Dowding RJ. Nanocrystalline Tungsten
Powder Synthesis Using High Energy Milling. Proc. 2008 Intl Conf on W, Refract &
pressed density, sintering results in significant microstructural Hardmetals VII, Publ MPIF, 5-35 – 5-48.
coarsening in tungsten powder compacts. For both the heating [8] Malewar R, Kumar KS, Murty BS, Sarma B, Pabi SK. On sinterability of
modes, the as-sintered compacts attain well-developed polyhedral nanostructured W produced by high-energy ball milling. J Mater Res 2007;22
(5):1200–6.
grains. The average tungsten grain size was measured using intercept [9] Johnson JL. Progress in Processing Nanoscale Refractory and Hardmetal Powders.
method and is summarized in Table 1. From Fig. 4 and Table 1, it is Proc. 2008 Intl Conf on W, Refract & Hardmetals VII, Publ MPIF, 5-57 – 5-71.
evident that varying heating modes have not significantly influenced [10] Wang H, Fang Z, Siddle D. Study of size-dependent sintering behavior of tungsten
powders. Tungsten, Refractory & Hardmetals VII. Princeton: Metal Powder
the microstructures.
Industries Federation; 2008. 5-72 – 5-77.
Table 1 also compares the effect of heating mode on the bulk [11] Oda1 E, Ameyama K, Yamaguchi S. Fabrication of nano grain tungsten compact by
hardness of the sintered tungsten compacts. As expected, the mechanical milling process and its high temperature properties. Mater. Sci.
microwave sintered compacts have higher hardness. This can be Forum; 2006. 503-504, 573-578.
[12] Jain M, Skandan G, Martin K, Cho K, Klotz B, Dowding R, et al. Microwave sintering:
attributed to the higher density and a more refined microstructure of a new approach to fine-grain tungsten-II. Int J Powder Metall 2006;42(2):53–7.
the tungsten grains in case of microwave sintering. The hardness of W [13] Jain M, Skandan G, Martin K, Cho K, Klotz B, Dowding R, et al. Microwave sintering:
compact achieved through microwave sintering in this study is higher a new approach to fine-grain tungsten-I. Int J Powder Metall 2006;42(2):45–50.
[14] Prabhu G, Charkraborty A, Sarma B. Microwave sintering of tungsten. Int J Refract
than those reported earlier by Upadhyaya et al. [15] and Kim et al. Hard Met 2009;27:545–8.
[28]. The hardness results achieved in microwave sintered W [15] Upadhyaya A, Tiwari SK, Mishra P. Microwave sintering of W–Ni–Fe alloy. Scripta
compacts in the present study is 30% higher than the best values Mater 2007;56:5–8.
[16] Luo SD, Yi JH, Guo YL, Peng YD, Li LY, Ran JM. Microwave sintering W–Cu
(303VHN) achieved by Prabhu et al. [14] by microwave sintering composites: analyses of densification and microstructural homogenization. J
tungsten compact at 1800 °C. Alloys Comp 2009;473:L5–9.
[17] Zhou C, Yi J, Luo S, Peng Y, Li L, Chen G. Effect of heating rate on the microwave
sintered W–Ni–Fe heavy alloys. J Alloys Comp 2009;482:L6–8.
4. Conclusions [18] Mondal A, Agrawal D, Upadhyaya A. Microwave Sintering of Refractory Metals/
Alloys: W, Mo, Re, W–Cu, W–Ni–Cu and W–Ni–Fe Alloys. J. Microwave Power
Electromagnetic Energy in press;44.
In this study, tungsten powder compact was successfully consol- [19] Mondal A, Agrawal D, Upadhyaya A. Microwave and conventional sintering of
idated through microwave sintering. It is interesting to note that premixed and prealloyed tungsten heavy alloys. Mater Sci Tech 2008:2502–15.
tungsten compact strongly couples with microwaves and gets heated [20] Mondal A, Upadhyaya A, Agrawal D. Microwave sintering of W–18Cu and W–7Ni–
3Cu alloys. J Microw Power Electromagn Energy 2009;43(1):11–6.
very rapidly. This result in the drastic reduction in the sintering cycle [21] Mondal A, “Modeling of Microwave Heating of Particulate Metals and its
time over the conventional process: up to as high as 90%. Despite Application in Sintering of Tungsten-Based Alloys”, Ph.D. Thesis, The Indian
being heated at high heating rates, none of the tungsten powder Institute of Technology, Kanpur, submitted.
[22] Brandes EA, Brook GB. Smithells metals reference book7th ed. . Oxford, UK:
compacts displayed any distortion or cracking during microwave Butterworth Heinemannn; 2000.
heating. This underscores the volumetric heating aspects of micro- [23] Groza JR. Sintering of nanocrystalline powders. Int J Powder Metall 1999;35:59.
waves. Moreover, due to the low thermal mass, microwave heating [24] Hahn H. Microstructure and properties of nanostructured oxides. Nanostructured
Mater 1993;2:251.
was less energy (power) consuming as compared to the equivalent [25] Averbach RS, Zhu H, Tao R, Hofler HJ. In: Bourell DL, editor. Sintering of
volume of material being consolidated conventionally. Another nanocrystalline materials: experiments and computer simulations. synthesis and
common observation is that the microwave sintered compacts processing of nanocrystalline powder. Warrendale, PA: TMS; 1996. p. 203–16.
[26] Bourell L, Groza JR. Consolidation of ultra fine and nanocrystalline powder. ASM
undergo more densification as compared to their conventionally
handbook, vol. 7. Metals Park, OH: American Society of Metals; 1998. p. 583–9.
processed counterparts. Because of the rapid heating rates achieved in [27] Groza JR. Nanocrystalline powder consolidation methods. In: Koch CC, editor.
microwave furnace, the microstructure coarsening was relatively Nanostructured materials—processing, properties and potential applications. New
lower. Consequently, as compared to conventional sintering, the York, NY: Noyes; 2002. p. 115–78.
[28] Kim YM, Hong MH, Lee SH, Kim EP, Lee S, Noh JW. The effect of yttrium oxide on
mechanical properties of the microwave sintered compacts were the sintering behavior and hardness of tungsten. Met Mater Int 2006;12(3):
higher. 245–8.
Comparative study of densification and
microstructural development in W–18Cu
composites using microwave and
conventional heating
A. Mondal*1, A. Upadhyaya2 and D. Agrawal3
Tungsten based composites such as W–Cu have been widely used as electrical contacts,
especially in heavy duty applications and as spark erosion electrodes. The lack of solubility
between tungsten and copper makes it very difficult to achieve full densification through liquid
phase sintering. Higher sintering temperatures or longer holding times always help to improve the
densification but Cu may leach out from the skeleton which leads to Cu segregation and results in
non-homogeneous microstructure and poor product performance. Microwave heating has been
increasingly gaining popularity in the field of sintering of particulate materials. As compared to
conventional heating, microwave heating is more rapid resulting in substantial reduction in the
overall sintering time. In addition to the energy efficiency, the faster heating rate achieved in
microwave furnaces minimises microstructural coarsening and improves homogeneity. This study
examines the effect of heating mode (conventional and microwave) and temperature on the
consolidation of specially prepared commercial W–Cu powder. Near theoretical density has been
achieved under optimum conditions in microwave sintering. The bulk hardness and electrical
conductivity of the samples sintered by two methods have been determined and the data
compared.
Keywords: Liquid phase sintering, Microwave sintering, Powder consolidation, Synthesis and processing

Introduction through LPS, sufficient amount of melt and its


homogeneous distribution is required.3 In the case of
Tungsten–copper alloys are one of the most important W–Cu compacts, while high compaction pressures
alloys of tungsten which have widespread use as increase the green density, it results in increased particle
electrical contact materials, in strategic applications contacts that limit particle rearrangement during LPS.4
(e.g. shape chargers) and for thermal management Another approach for ensuring densification in the
applications.1,2 On account of the refractoriness of the W–Cu system is through the addition of transition metal
major phase (Wmp: 3420uC), these alloys are usually additives that activate the sintering kinetics.5–7 The
processed through liquid phase sintering. There are two addition of small quantities of transition elements (such
ways used to form W–Cu composites: copper infiltration as Ni, Co and Fe), activates tungsten skeletal sintering
of a porous tungsten skeleton and liquid phase sintering and has a positive impact on the sinterability of the W–
(LPS) of mixed phase compacts. For economical Cu. Such activators remain segregated at the grain
reasons, shape, complexity and large scale production boundaries and thereby result in poor mechanical
consideration, the latter technique is the only widely
properties.8 Higher sintering temperature or longer
adopted fabrication routes for W–Cu alloys. The lack of
holding time always helps to improve the densification
solubility between tungsten and copper makes it very
but copper may leach out from the skeleton which leads
difficult to achieve full densification through liquid
to non-homogeneous microstructure.9 Co-reduction of
phase sintering. For ensuring complete densification
oxides or other thermochemical processes are also
efficient methods to improve the densification of this
1
NTPC Energy Technologies and Research Alliance (NETRA), Greater material.3,5,10 Novel sintering techniques have also been
Noida-201306, INDIA
2
explored to enhance W–Cu composite densifications
Department of Materials Science and Engineering Indian Institute of
Technology, Kanpur 208016, INDIA
which include resistance sintering, plasma spraying and
3
Microwave Processing and Engineering Center, Materials Research laser sintering.11,12
Institute, The Pennsylvania State University, PA 16802, USA Microwave sintering is a well known technique in the
*Corresponding author, email avijit.mondal@gmail.com field of sintering of particulate ceramic materials. But for

ß W. S. Maney & Son Ltd. 2010


Received 14 May 2010; accepted 25 May 2010
DOI 10.1179/143307510X12820854748638 Materials Research Innovations 2010 VOL 14 NO 5 355
Mondal et al. Densification and microstructural development in W–18Cu composites

the metallic material it was not applied earlier because of


the fact that bulk metal reflects microwave. But recently
it has been shown that metals can also be coupled with
microwave and get heated provided they are in powder
form.13 This has resulted in wide spread interest in
sintering of many metallic compacts in microwave field.
As compared to conventional heating, microwave
heating is more rapid which results in reducing the
overall sintering time. In addition to the cost efficiency,
the faster heating rate achieved in microwave techniques
1 Scanning electron micrograph of as received W–18Cu
minimises microstructural coarsening. For powder
powder: a and b are at different magnifications
compacts sintering is usually associated with densifica-
tion as well as concomitant microstructural coarsening.
applied uniaxially in a 50 t hydraulic press (model
If the latter is restricted the mass atomic transport is
CTM-50, FIE, Ichalkaranji, India) with floating die,
enhanced due to the availability of more grain bound-
using zinc stearate as a die wall lubricant.
aries. Recent review reports14 on various systems have
All the samples were compressed at pressures of
exhibited higher sintered density and better mechanical
200 MPa. The as pressed green compact density varied
properties in microwave sintered compacts as compared
from 52 to 54% of the theoretical density of the
to their conventionally sintered counterparts.
composites, where the theoretical density rT was
The application of microwave energy in consolidating
calculated using the inverse rule of mixtures as follows
tungsten powders was first studied by Jain et al.15 A
comprehensive study on microwave sintering of tungsten 1 X
N
wi
and its alloys were also conducted by several research- ~ (2)
rT i~1 ri
ers.16–20 In a very recent report, Luo et al.21 studied the
microwave sintering of W–Cu alloys. However, their where N is the number of elements in the mixture, wi is
assessment was qualitative, and the comparative analysis the weight fraction of the ith component and ri is the
was lacking in terms of properties concerned. theoretical density of the ith element. To study the
This study investigated the effect of heating mode as densification behaviour, the green (as pressed) compacts
well as temperature on the densification, microstructure were sintered using conventional and microwave fur-
and properties of specially prepared commercial W– nace. The conventional sintering of green compacts was
18Cu powder compacts. Dilatometry measurements carried in a MoSi2 heated horizontal tubular sintering
have also been used to study the in situ linear shrinkage furnace (model OKAY 70T-7, Bysakh, Kolkata, India)
behaviour as a function of temperature. at a constant heating rate of 5uC min21. To ensure
uniform temperature distribution during heating, inter-
Experimental mittent isothermal hold for 30 min was provided at 500
and 900uC. The sintering temperatures selected for the
Table 1 summarises the characteristics of the as received solid state and LPS were 1000, 1200, 1300 and 1400uC
W–18Cu power used in this investigation. The spatially respectively. Sintering was carried out for 60 and 30 min
prepared powder was supplied by Osram Sylvania, in H2 atmosphere with dew point 235uC. Microwave
Towanda, PA, USA. Particle size analysis of the as sintering of the green compacts was carried out using a
received powder was performed using Malvern multimode cavity 2?45 GHz, 6 kW commercial micro-
Instruments particle size analyser which works on the wave furnace (model RC/20SE, Amana Radarange,
principle of laser diffraction. The surface area of the as Benton Harbor, MI, USA).
received powder was measured by nitrogen adsorptions Unlike a conventional furnace, the temperature of the
method using a Brunauer–Emmett–Teller (BET) appa- samples inside a microwave furnace cannot be mon-
ratus. The BET equivalent spherical particle diameter itored by using a thermocouple. The presence of
DBET was calculated for the powders using equation (1) thermocouples can locally distort the electromagnetic
6 field and can even lead to measurement errors. The
DBET ~ (1) temperature of the sample was monitored using an
rp A
optical pyrometer (Marathon Series, Raytek, Santa
where A is the specific BET surface area of the powder in Cruz, CA, USA) with the circular crosswire focused
m2 g21 and rp is the pycnometer density of the powder on the sample cross-section. Further details of the
in g cm23 and DBET is in mm. experimental set-up of microwave sintering are
Figure 1 is a scanning electron micrograph of as described elsewhere.22
received powder. Powders were pressed in a die of The sintered density was obtained by both dimen-
1?6 cm inner diameter to make the green compacts of sional measurements as well as Archimedes’ density
approximately 0?6–0?8 cm in height. The pressure was measurement technique. To take into account the

Table 1 Powder characteristics of as received W–18Cu powder

Powder size, mm
Surface BET equivalent
Property Processing technique Powder shape D10 D50 D90 area, m2 g21 particle size, mm

W–18Cu Chemical synthesis of Spherical 1.74 8.23 33.19 0.81 0.47


powder W–Cu oxides

Materials Research Innovations 2010 VOL 14 NO 5 356


Mondal et al. Densification and microstructural development in W–18Cu composites

graphs of as polished samples were obtained by a


scanning electron microscope (model FEI Quanta, FEI
Company, Eindhoven, The Netherlands) in the second-
ary electron and backscattered mode. Phase determina-
tion and phase evolution, if any were studied for all the
samples with the help of X-ray diffractometer (model
ISO Debyeflex-2002, Rich Seifert & Co., Ahrensburg,
Germany) with a scan rate of 3u min21. Bulk hardness
measurements were performed on polished surfaces of
sintered cylindrical compacts at a load of 5 kg using a
Vickers hardness tester (model V100-C1, Leco, Tokyo,
Japan). The load was applied for 30 s. The Vickers bulk
hardness value for each sample was an average of 10
readings taken at random locations throughout the
sample. The electrical conductivity measurements of the
2 Thermal profile of W–18Cu compact heated in radia- sintered samples were carried out using digital con-
tively heated and 2?45 GHz microwave furnace ductivity meter (model 757, TechnoFour, Pune, India).

influence of the initial as pressed density, the compact Results and discussion
sinterability was also expressed in terms of densification Heating profiles
parameter which is calculated as follows
The W–18Cu powder compacts were sintered for 30 min
Densification parameter~ at temperatures corresponding to solid state (1000uC)
and liquid phase (1200, 1300 and 1400uC) sintering.
sintered density{green density Figure 2 compares the thermal profiles of the compacts
theoretical density{green density (3) heated in conventional and microwave furnace. It is
The sintered samples were wet polished in a manual interesting to note that W–18Cu compact couples with
polisher (model Major, Struers, Ballerup, Denmark) microwaves very efficiently and gets heated up rapidly.
using a series of 6, 3 and 1 mm diamond paste, followed The overall heating rate in microwave furnace was
by cloth polishing using a suspension of 0?04 mm y22uC min21; taking into consideration the slow
colloidal SiO2 suspension. The scanning electron micro- heating rate (5uC min21) and isothermal holds at
intermittent temperatures in conventional furnace, there
is y80% reduction in the overall process time during
sintering of W–18Cu compacts in microwave furnace.
Figure 3 shows the photograph of the conventionally
and microwave sintered W–18Cu compacts. From the
figure, it is evident that in both cases the sintered W–
18Cu compacts do not show any macrocracking or
dimensional distortion.
Densification response of W–18Cu
Figure 4a and b shows the effect of sintering tempera-
ture and heating mode on the sintered density and
densification parameter respectively. Since all compacts
3 Photograph of W–18Cu compacts sintered at 1200uC in were pressed to the same green density, hence, the
a conventional and b microwave furnace densification parameter response is similar to sintered

4 Effect of heating mode and sintering temperature on a sintered density and b densification parameter of W–18Cu alloy

Materials Research Innovations 2010 VOL 14 NO 5 357


Mondal et al. Densification and microstructural development in W–18Cu composites

density. For temperatures corresponding to LPS, both


microwave and conventional sintering methods result in
similar (nearly 100%) densification. In fact, at 1200uC
itself the compact attains full densification which
underscores the efficacy of capillary induced rearrange-
ment. It is, however, remarkable to observe the positive
influence of microwave heating on compacts consoli-
dated in solid state at 1000uC. These results are
contradictory to those reported by Ryu et al.23 who
showed that the overall densification in W–Cu alloys
consolidated through solid state conventional sintering
is higher for compacts heated at slower heating rates. As
compared to conventional heating, microwave heating
rate is nearly five times higher. Thus, microwave heating
not only lowers significantly the processing time but also
enhances the sintering kinetics.
Figure 5 compares the effect of heating mode and
sintering temperature on the axial and radial shrinkage
5 Effect of heating mode and sintering temperature on of the cylindrical compacts. Note that irrespective of the
axial and radial shrinkage of W–18Cu alloy sintering temperature and the heating mode dimensional
changes in the W–18Cu compacts are isotropic in
nature.
The as sintered W–18Cu alloys were metallographi-
cally prepared for microstructural observation. Figure 6a
and b compares the microstructures of the W–18Cu
alloys solid state sintered (1000uC) in a conventional and
microwave furnace respectively. Figure 7a–c compares
the effect of temperature on the microstructure of W–Cu
compacts consolidated in conventional and microwave
furnace through LPS. For both the heating modes,
higher sintering temperatures are expected to lead to
6 Representative microstructure of W–18Cu compacts
solid state sintered at 1000uC in a conventional and
microstructural coarsening. In the case of conventionally
b microwave furnace
sintered W–18Cu compacts, microstructural coarsening
is distinctly observed. At 1400uC, the tungsten grains
appear in well defined spheroidal form. Figure 8 shows
the elemental distribution in W–Cu compacts sintered at
1400uC. It is to be noted that the distribution of W and
Cu phases is quite distinct corresponding to granular
structure and matrix region respectively in the sintered
body, while no intermixing was observed. This confirms
lack of mutual solubility between W and Cu. Unlike
systems such as W–Ni–Fe and W–Ni–Cu alloys, the lack
of solid solubility obviates solution reprecipitation in W–
Cu alloys during sintering.24,25 Hence, grain coarsening
in W–Cu alloys is restricted and coalescence is the sole
contributor to it.26 As compared to conventional
sintering, shorter processing time in microwave sintering
restricts microstructural coarsening still further.

X-ray diffraction (XRD) analysis


Figure 9 compares the XRD patterns of the composition
in both the heating modes at different temperatures.
From XRD patterns of both conventionally and
microwave sintered samples, no difference as far as
phase evolution is concerned was observed. However, it
is interesting to note that although the sample size and
other parameters were constant during the experiment
still relative intensity of the X-ray peaks was slightly
lower in microwave sintered samples than conventional
one. This difference dominates at higher temperature.
There is no plausible explanation for this observation
7 Scanning electron micrographs of W–18Cu alloys sin- yet. One possible explanation may be that in microwave
tered at a 1200uC, b 1300uC and c 1400uC in conven- the W phase may be undergoing a partial decrystallisa-
tional (left) and microwave (right) furnace tion as had been observed in a single mode microwave

Materials Research Innovations 2010 VOL 14 NO 5 358


Mondal et al. Densification and microstructural development in W–18Cu composites

10 Effect of heating mode and sintering temperature on


bulk hardness of W–18Cu alloys

alloys. The variation in the hardness closely is related to


b the variation in the sintered porosity. For both the
heating modes, liquid phase sintered compacts have
significantly higher hardness as compared to those
consolidated at 1000uC through solid state sintering.
The size of the tungsten grains also seems to influence
the bulk hardness. In the case of W–18Cu alloys
conventionally processed through LPS, the hardness
decreases as the sintering temperature increases from
1200 to 1400uC. This can be correlated to the micro-
structural coarsening at higher temperatures (Fig. 7).
In the case of W–Cu compacts processed in micro-
c wave furnace, better binder homogeneity achieved at
8 a scanning electron micrograph of W–18Cu alloy liquid higher temperature, causes a gradual improvement of
phase sintered at 1400uC, b and c corresponding ele- the bulk hardness as the sintering temperature increases.
mental maps Figure 11 compares the effect of heating mode and
sintering temperature on the electrical conductivity of
W–18Cu compacts. The trend in electrical conductivity
H-field processing of certain materials such as ferrites, variation is similar to that observed for hardness
Si, etc.27,28 (Fig. 10). In general, microwave sintered compacts
exhibit improved conductivity. For both the heating
Bulk hardness and electrical conductivity modes, as compared to solid state sintering, LPS
Figure 10 compares the effect of heating mode and resulted in significant improvement in the conductivity.
sintering temperature on the bulk hardness of W–18Cu This can be attributed to greater densification and more

11 Effect of sintering temperature on electrical conductiv-


9 X-ray diffraction pattern of conventionally and micro- ity of W–18Cu compacts consolidated in conventional
wave sintered W–18Cu alloys (radiatively heated) and microwave furnace

Materials Research Innovations 2010 VOL 14 NO 5 359


Mondal et al. Densification and microstructural development in W–18Cu composites

uniform binder distribution in liquid phase sintered 7. J. L. Johnson and R. M. German: ‘Advances in powder metallurgy
and particulate materials’, Vol. 4, 201–213; 1993, Princeton, NJ,
compacts.
MPIF.
8. A. Upadhyaya and R. M. German: Int. J. Powder Metall., 1998, 34,
Conclusion 43–55.
9. J. L. Johnson, J. J. Brezovsky and R. M. German: Metall. Mater.
Compacts of W–18Cu alloys were successfully sintered in Trans. A, 2005, 36A, 2807–2814.
a microwave furnace for temperatures corresponding to 10. O. Ozer, J. M. Missiaen, S. Lay and R. Mitteau: Mater. Sci. Eng.
solid and liquid phase sintering in relatively much shorter A, 2007, A460–A461, 525–532.
time than in the conventional heating method. In spite of 11. Z. J. Zhou, J. Du, S. X. Song, Z. H. Zhong and C. C. Ge: J. Alloys
Compd, 2007, 428, 146–150.
higher heating rate in microwave sintering, no micro- or 12. G. Pintsuk, S. E. Brünings, J. E. Döring, J. Linke, I. Smid and
macrocracks in the sintered samples were observed which L. Xue: Fusion Eng. Des., 2003, 66–68, 237–241.
essentially supports the volumetric heating nature of 13. R. Roy, D. K. Agrawal, J. P. Cheng and S. Gedevanishvili: Nature,
microwave sintering. Microwave sintering lowers the 1999, 399, 668–670.
14. M. Oghbaei and O. Mirzaee: J. Alloys Compd, 2010, 494, 175–189.
sintering temperature to achieve the optimum densifica-
15. M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding,
tion as well as optimum mechanical properties. Finer D. Kapoor, D. Agarwal and J. Chang: Int. J. Powder Metall., 2006,
particle size and more homogeneous distribution of 42, 45–50.
tungsten in the microstructure can be obtained by using 16. G. Prabhu, A. Charkraborty and B. Sarma: Int. J. Refract. Hard
microwave energy for the consolidation of W–18Cu Met., 2009, 27, 545–548.
17. C. Zhou, J. Yi, S. Luo, Y. Peng, L. Li and G. Chen: J. Alloys
composites leading to higher mechanical properties. Compd, 2009, 482, L6–L8.
18. A. Upadhyaya, S. K. Tiwari and P. Mishra: Scr. Mater., 2007, 56,
Acknowledgements 5–8.
19. A. Mondal, D. Agrawal and A. Upadhyaya: J. Microw. Power
The authors would like to acknowledge Osram Sylvania, Electromagn. Energy, 2010, 44, to be published.
Towanda, PA, USA, for supplying the powder used for 20. A. Mondal: ‘Modeling of microwave heating of particulate metals
this investigation. Partial support from Indo–US Science and its application in sintering of tungsten–based alloys’, PhD
thesis, Indian Institute of Technology, Kanpur, India, 2009.
and Technology Forum (IUSSTF), New Delhi, India is 21. S. D. Luo, J. H. Yi, Y. L. Guo, Y. D. Peng, L. Y. Li and J. M. Ran:
gratefully acknowledged. J. Alloys Compd, 2009, 473, L5–L9.
22. A. Mondal, D. Agrawal and A. Upadhyaya: J. Microw. Power
References Electromagn. Energy, 2009, 43, 5–10.
23. S. S. Ryu, Y. D. Kim and I. H. Moon: J. Alloys Compd, 2002, 335,
1. A. Upadhyaya: Mater. Chem. Phys., 2001, 67, 101–110. 233–240.
2. A. Upadhyaya: Trans. Indian Inst. Met., 2002, 55, 51–69. 24. F. V. Lenel: Trans AIME, 1948, 175, 878–896.
3. J. L. Johnson, J. J. Brezovsky and R. M. German: Metall. Mater. 25. W. J. Huppmann: Z. Metallkd, 1979, 70, 792–797.
Trans. A, 2005, 36A, 1157–1165. 26. E. G. Zukas, P. S. Z. Rogers and R. S. Rogers: Z. Metallkd, 1976,
4. J. L. Johnson and R. M. German: Metall. Mater. Trans. B, 1996, 67, 591–595.
27B, 901–909. 27. R. Roy, Y. Fang, J. Cheng and D. K. Agrawal: J. Amer. Ceram.
5. D. G. Kim, G. S. Kim, M. J. Suk, S. T. Oh and Y. D. Kim: Scr. Soc., 2005, 88, 1640–1642.
Mater., 2004, 51, 677–681. 28. R. Peelamedu, R. Roy, L. Hurtt, D. Agrawal, A. W. Fliflet,
6. V. Gauthier, F. Robaut, A. Upadhyaya and C. H. Alliberta: D. Lewis, III and R. W. Bruce: Mater. Chem. Phys., 2004, 88, 119–
J. Alloys Compd, 2003, 361, 222–226. 129.

Materials Research Innovations 2010 VOL 14 NO 5 360


Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010, pp. 28-44
A Publication of the International Microwave Power Institute

Microwave Sintering of Refractory


Metals/alloys: W, Mo, Re, W-Cu,
W-Ni-Cu and W-Ni-Fe Alloys
Avijit Mondal1, Dinesh Agrawal2, Anish Upadhyaya1
1
Department of Materials & Metallurgical Engineering
Indian Institute of Technology, Kanpur 208016, INDIA

2
Materials Research Institute
The Pennsylvania State University, University Park, PA 16802, USA

Received: August 5, 2009


Accepted: January 28, 2010

ABSTRACT
Refractory metals and alloys are well known for their high mechanical properties which
make them useful for wide range of high temperature applications. However, owing to the
refractoriness of these metals and alloys, it is very difficult to consolidate them under moderate
conditions. Conventional P/M processing is a viable sintering technique for these refractory
metals. One of the constraints in conventional sintering is long residence time which results in
undesirable microstructural coarsening. This problem gets further aggravated when using smaller
(submicron and nano) precursor powder sizes. Furthermore, conventional heating is mostly
radiative, which leads to non-uniform heating in large components. This review article describes
recent research findings about how these refractory metals and alloys (W, Mo, Re, W-Cu, W-Ni-
Cu and W-Ni-Fe) have been successfully consolidated using microwave sintering. A comparative
study with conventional data has been made. In most cases, microwave sintering resulted in an
overall reduction of sintering time of up to 80%. This sintering time reduction prevents grain
growth substantially providing finer microstructure and as a result better mechanical properties
have been observed.

KEYWORDS: Microwave sintering; Refractory metals/alloys; Microstrutures

INTRODUCTION
Refractory metals are known for their very high melting temperatures. Most refractory
metals used for various applications are tungsten with fusion point of 3420°C, molybdenum of
2620°C and rhenium of 3180°C. Because of their high melting point, most refractory metals and
alloys are consolidated through powder metallurgy (P/M) techniques, though for some specific
applications mechanical alloying and infiltration technique are also employed. Most commonly,
liquid phase sintering (LPS) is used for consolidating tungsten based alloys such as W-Cu, W-
Ni-Cu and W-Ni-Fe compositions [Upadhyaya, 2001]. The LPS offers an advantage of relatively
lower sintering temperature, enhanced densification, microstructural homogenization and near
theoretical density.

International Microwave Power Institute


28
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

These refractory metals and alloys penetrators in ordnance industry. For most
offer a wide range of properties which make of the applications, near theoretical density,
them advantageous for high temperature dimensional stability, higher hardness,
applications over other materials. Sintered toughness and very high ductility are
tungsten is an excellent material for many important.
applications such as lightings, heating, These metals and alloys in conventional
aerospace, electronic, sports and military furnace are sintered at very high temperatures
uses due to its high melting point, high (1500-2000°C) and to avoid thermal shock at
density of 19.3 g/cm3, high hardness of 9.75 a slow heating rate (<10°C/min) and with an
GPa, moderate elastic modulus of 407 GPa, isothermal hold at intermittent temperatures.
low coefficient of thermal expansion, good The high sintering temperature results in
thermal conductivity, and low vapor pressure. significant microstructural coarsening in the
Rhenium metal is interesting from a number sintered material, leading to the degradation
of standpoints. It is only second to tungsten, of mechanical properties. This problem is
among the metallic elements, in melting further aggravated when the initial powder
point. Its density of 21.0 g/cm3 is higher size is extremely fine. Hence, it is envisaged
than that of tungsten. Annealed material has that a sintering method that provides a rapid
exhibited tensile strengths of about 120,000 heating rate, lower sintering temopertaure
p.s.i. with 25% ductility at room temperature, and duration would mitigate this problem.
and it is somewhat harder and more resistant One of the techniques to achieve rapid
to abrasion than tungsten. Other properties, and relatively uniform sintering is through
such as its corrosion resistance and electrical microwaves [Rao et al., 1995; Clark and
properties make it promising for incandescent Sutton, 1996; Agrawal, 1998].
lamp filaments and electrical contacts. Microwave sintering in the recent
Molybdenum is a typical transition metal times has emerged as an innovative technique
element having a high melting point, high for high temperature material processing.
mechanical strength, and high modulus of Microwave assisted synthesis is generally
elasticity. Most of the applications for pure faster, cleaner and more economical than
molybdenum metal and its alloys involve the conventional methods. The possibility of
as electrodes for electrically heated glass ceramics processing by microwave heating
furnaces and forehearths, nuclear energy was first discussed over 50 years ago by Von
applications, missile and aircraft parts, Hippel [Von Hippel, 1954], and experimental
thermocouple sheaths, flame and corrosion studies started in the middle of the 1960s by
resistant coatings for other metals, and as an Tinga and co-authors [Tinga et al., 1968; Tinga
alloying agent in steel. and Edwards, 1968]. Since then variety of
An important class of tungsten based materials such as carbides, nitrides, complex
material is tungsten heavy alloys (WHA). A oxides, silicides, zeolieties apatite, etc. have
typical tungsten heavy alloy contains 60 to been synthesized using microwaves [Bykov et.
98 wt. % tungsten. The balance is generally al., 2001; Booske et al., 1997; Sutton, 1992;
a mixture of relatively low melting transition Agrawal, 1999; Agrawal et al., 2001; Rodiger
elements, such as nickel, iron, copper, and et al., 1998; Agrawal et al., 2000].
cobalt. Due to their unique combination In 1999, for the first time Roy et al.
of properties WHAs have a wide range [Roy et al., 1999] reported that a porous
of applications, such as radiation-shield, powder metal compact could be heated
counter-balanced weights etc. and sintered in a microwave field. Their
Most important of all applications is work added a new dimension towards the
its potential to be a replacement of depleted application of microwave energy for high
uranium within kinetic energy anti armor temperature material processing. The

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


29
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

results of many investigations in microwave 1996; Agrawal, 1998; Von Hippel, 1954; Tinga
sintering and joining of ceramics, ceramic- et al., 1968; Tinga and Edwards, 1968; Bykov
metal composites, metals and alloys have et. al., 2001; Booske et al., 1997; Sutton,
been reported [Bykov et. al., 2001; Booske 1992; Agrawal, 1999; Agrawal et al., 2001;
et al., 1997; Sutton, 1992; Agrawal, 1999; Rodiger et al., 1998; Agrawal et al., 2000].
Agrawal et al., 2001; Rodiger et al., 1998; Microwave heating in metals is different from
Agrawal et al., 2000; Anklekar et al., 2001; that observed in dielectric materials (mostly
Saitou, 2006; Luo, J. et al., 2004; Anklekar ceramics). Being good conductors, no internal
et al., 2005; Takayama et. al., 2006; Rybakov electrical field is induced in metals.
et al., 2006; Mishra et al., 2006; Mondal The induced electrical charge remains
et al., 2008; Mondal et al., 2009]. In the at the surface of a bulk metallic sample. As a
majority of the papers the authors claim consequence, bulk metals reflect microwaves
acceleration of microwave driven processes at room temperature; hence no bulk
as compared with the processes performed absorption (heating) occurs, particularly, at
using conventional heating. The acceleration temperatures below 500°C. According to the
commonly manifests itself as a reduction in the Faraday’s effect in a conductive material, a
densification time of the powder compacts, varying magnetic field generates an electric
which is often accompanied by a decrease in field that gives rise to eddy currents and
the temperature of sintering. High rates of subsequently resistive losses. Additionally
volumetric heating in microwave sintering, during sintering of particulate metal
not limited by thermal diffusion, prevent compact, each individual powder particle in
recrystallization grain growth and result in the compact is surrounded by a dielectric
a finer and more uniform microstructure in oxide layer. The presence of such dielectric
the sintered bulk materials. It is well known “shell” on the powder particle prevents the
that a fine, homogeneous, and fault-free connectivity percolation between the particles
microstructure is a necessary prerequisite for and increases significantly electromagnetic
enhanced material performance. Similarly, power that can be absorbed by the compact.
a decrease in the duration of the high- In general, the skin depth is relatively small
temperature stage leads to reduced grain in metals, since in the microwave regime,
growth as a result, to the higher mechanical the particle sizes are much smaller than the
strength in the final product. wavelength of microwave radiation; the field
There are different mechanisms across the particle are uniform and causes
by which microwaves can couple to a volumetric heating. However, for relatively
material and a whole host of ways that the coarse particle (>100µm), the heating may be
microwave energy is subsequently absorbed conductive from outside to the interior of the
by the system. The main loss mechanisms are powder. Recent study confirms that magnetic
electric, conduction (eddy current), hysteresis heating plays most important role for metallic
and resonance (domain wall and electron spin materials. Quite contrary to the fact that
(FMR)). It is often difficult to ascertain which electric heating predominates for dielectric
loss mechanism, or combination of mechanisms materials. The detailed microwave absorption
is occurring for a particular material in given mechanism for metallic material can be found
conditions. The different mechanisms do in the reference already mentioned.
however have different dependencies on From the above discussion it is quite
certain properties such as sample type and clear that microwave processing has many
microstructure, frequency and temperature. advantages over conventional methods,
A brief description of these different loss especially for sintering applications. The
mechanisms for ceramic material can be various advantages are: time and energy
found in the reference [Clark and Sutton, saving, rapid heating rates, considerably

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


30
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

reduced processing cycle time and crystalline tungsten (free of grain boundaries)
temperature, substantial enhancement remains ductile down to at least 20K. Any
in reactivity and diffusion kinetics, fine plastic deformation decreases the transition
homogeneous microstructures and improved temperature by fining the structure [Lassner
mechanical properties which lead to better and Schubert, 1999]. The presence of small
product performance. Microwave-matter amount of interstitial impurities such as O, C,
interaction and anisothermal situations often N and H2 has a detrimental effect on DBTT.
produce better quality and new materials that The solubility for such elements in
are normally not possible with conventional tungsten at room temperature is much lower
methods. These characteristics advantages when compare to high temperature. So during
of microwave sintering can be exploited to cooling segregation of such elements at the
overcome the processing difficulties of the grain boundary significantly weakens the grain
refractory metals and alloys. boundary strength. The ductilizing effect
This review article describes recent of Re in tungsten have gained outstanding
research findings about how certain refractory importance in this regard. Addition of Re can
metals and alloys (W, Mo, Re, W-Cu, W-Ni- cause the transformation temperature to fall
Cu and W-Ni-Fe) have been successfully below room temperature, even for slightly
consolidated using microwave sintering and deformed products. Many studies have shown
a comparative study with conventional data that sintering temperature is related to the
has been made. In most cases, microwave powder size, when the size is in nano-scale,
sintering resulted in an overall reduction of the sintering temperature can be decreased
sintering time of up to 80%. This sintering time up to several hundred degrees. The reduction
reduction prevents grain growth providing of sintering temperature for nano tungsten
finer microstructure and as a result better has been reported by several researchers
mechanical properties have been observed. [Sarma and Pabi, 2007; Oda et al., 2006;
Bose et al., 2008; Engleman et al., 2008;
CONSOLIDATION OF TUNGSTEN Johnson, 2008; Wang et al., 2008; Jain
Usually the consolidation of W powder et al., 2006; Jain et al., 2006]. The reported
by conventional heating is difficult and sintering temperature of nano sized tungsten
requires very high temperature (2200°C or produced by high energy mechanical milling
more) in electrical resistance sintering under was drastically decreased from conventional
hydrogen atmosphere. The requirement temperature of 2500°C to 1700°C [Malewar
of excessive high temperature and special et al., 2007]. Researchers have also shown
technique makes the process more expensive that pressure assisted process such as spark
and imparts a restriction in the sizes and plasma sintering [Oda et al., 2006], plasma
shapes of the sintered products. pressure compaction [Bose et al., 2008],
Low-temperature brittleness is the and hot isostatic processing (HIP) [Engleman
most crucial aspect in the manufacturing of et al., 2008] etc. helps in further reduction
pure tungsten metal. Therefore, in the past in the processing temperature. Selection of
much effort has been directed at lowering nano powders over microcrystalline powders
the ductile to- brittle transition temperature has certain distinct advantages. According to
(DBTT) and hence improving the fabricability the literature reported data for both metals
of the metal. and ceramic nano particles was found to start
The brittleness of polycrystalline densification at temperatures of 0.2–0.4 Tm
tungsten at low temperature is attributed to (Tm - melting temperature) compared to 0.5–
the weakness of the grain boundaries, which 0.8 Tm for the conventional powders.
leads to initiation of cracking in both wrought It is believed that in the case of
and recrystallized tungsten. Pure, single nanostructured powders the grain boundary

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


31
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

sliding, dislocation motion, grain rotation, Y2O3 as successful grain growth inhibitors.
and viscous flow can significantly contribute They also observed that the introduction of
to the enhanced sintering kinetics [Hahn, a secondary oxide (HfO2 and/or Y2O3) had a
1993; Averbach et al., 1996; Bourell and significant effect on the powder morphology
Groza, 1998; Groza, 1999; Groza, 2002]. and in reducing the primary particle size of
Generally sintering theories hold that the the as synthesized tungsten powders. The
sintering temperature is often correlated particle size was reduced from 350 nm to 80-
with melting point of the material. It is long 100nm, and the crystallite size was reduced
been know that the melting temperature from 48 nm to 25 nm with the addition of
of very fine particles decreases with the dopents [Jain et al., 2006].
size of the particles [Couchman and Jesser, Prabhu [Prabhu et al., 2008] has also
1977]. Therefore in addition to the faster investigated microwave sintering of pure
sintering kinetics, the faster densification in tungsten powder of as received grade and
nano structured material could be attributed tungsten powder activated by high energy
to the lower melting temperature of nano milling (HEM). Their study shows better
particles. sinterability of activated tungsten powder in
In addition to all the above mentioned compare to as received powder.
advantages, nano tungsten has two distinct Mondal [Mondal et al., 2009] has also
advantages over conventionally sized tungsten reported the similar kind of accelerated
powder. First, the thermal conductivity densification in microwave sintered tungsten
of nano tungsten is much lower than that sample. They reported that more than 95%
of conventionally sized tungsten powder. Th density of microwave sintered samples
Secondly, as in nano dimension grain size are at 1600°C for 30 min holding in hydrogen
close to dislocation length, no pinning or pile atmosphere.
up of dislocation occurs [Andrey et al., 2004]. Figure 1 describes typical thermal
The contribution of these two phenomenon profile used for their experiments in both
results in a material that exhibits a higher conventional as well as microwave heating
propensity of adiabatic shear, which is very mode.
important aspects of this material as a kinetic Figure 2 are the SEM micrograph
energy penetrator application is concerned. of both conventionally and microwave
Wang [Wang et al., 2008] has studied the size
dependent sintering behavior of tungsten
powder.
They observed that the starting
particle size before milling plays an important
role in sinterability of this material, lower
the size higher will be the sinterability.
Although, the final crystallite size after
milling is same in both the cases. The reason
for this difference could be attributed with
the difference in particle size after milling.
Mohit et al. [Jain et al., 2006] studied
the application of microwave energy in
consolidating pure tungsten powder. They
found 10 to 12% higher sintered density in
microwave sintering as compare to their
conventional counter part. In another study Figure. 1. Typical thermal profiles of sintered W using
they have reported the role of HfO2 and conventional and microwave furnaces [Mondal et al., 2009].

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


32
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Figure. 2. SEM micrographs of (left) conventional and (rigtj) microwave sintered W at 1600°C for 30 min in H2
atmosphere [Mondal et al., 2009].

Figure.3. Microwave sintering of complex shaped tungsten samples.

sintered samples. Figure 3 show some of an inert atmosphere (argon) or in a reducing


the commercial pure tungsten products that atmosphere (hydrogen) [Patrician et al.,
have been successfully sintered in microwave 1985]. High temperatures in the range of
furnace at Penn State’s Microwave Processing 2000°C are employed, resulting in densities
Center. of 90–95% of theoretical, depending upon the
sintering time. Huang [Huang and Hwang,
CONSOLIDATION OF MOLYBDENUM 2002] reported the sintering of molybdenum
Conventionally the sintering of using vacuum furnaces and obtained densities
molybdenum powder is conducted using a of 97 to 98.5% at a sintering temperature of
resistance or induction sintering furnace in 1750°C with times ranging from 10 to 40 h.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


33
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

This also results in abnormal grain growth. sintering times and temperatures produce
Microwave sintering of molybdenum metal bars of high yield and small grain size with
powder has been reported for the first time the same homogeneity attainable by electron
by Chillar [Chhillar et al., 2008]. In their beam or arc melting. Subsequent fabrication
work, the authors reported sintering of nano is performed by swaging, rolling, forging and
molybdenum powder to obtain submicron grain drawing with intermediate annealing.
size microstructure using microwave energy. Some of these operations may have to
As received Mo powder was agglomerated be conducted at elevated temperatures. The
with a mean agglomerate size of 1.6 um, optimum sequence of these operations varies
but equivalent surface area based on N2 for rhenium and it alloys and depends on the
adsorption suggests an average particle size final end form. Microwave sintering of rhenium
of 200 nm. Sintering was carried out using the pallet has been successfully conducted at
as received powder. Samples with densities as Penn State’s microwave processing center.
high as 98% of theoretical density (TD) were
obtained with limited grain growth in 5 min
of sintering time in microwaves, compared to
10–20 h in a conventional process.
The highlight of this research was
achieving 98%TD in 1 min at 1650°C with a
submicron grain size. Microwave sintering
result showed that near theoretical
densities can be obtained at much reduced
temperatures, and with much reduced
sintering times, as compared to conventional
sintering.
Conventionally sintered samples at
1400°C for 10 h resulted in 98%TD. However,
using microwave energy 99%TD could be
obtained at 1400°C in just 30 min. This
conclusively shows that microwave sintering
is much faster than conventional sintering.

CONSOLIDATION OF RHENIUM
Arc melting of rhenium in an inert
atmosphere or vacuum is possible but the
metal produced tends to have coarse grain
size and may have segregation of rhenium
oxides at the grain boundaries. These issues
are a problem for further fabrication of a
product and therefore powder metallurgy has
shown to mitigate some of these problems.
Rhenium powder is consolidated using pressure
techniques to a density of approximately 60%
of the theoretical density.
The pressed compacts are then pre-
sintered in a hydrogen atmosphere to facilitate Fig. 4: SEM micrographs of Re pallet in (above) as
handling before final sintering. Proper choice pressed and (below) microwave sintered at 2000°C for
of powder sizes, careful blending and adequate 10 min [Mondal et al., 2009].

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


34
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Relatively high sintered density in the order enhance W-Cu composite densification which
of 95% of theoretical has been achieved in includes resistance sintering, plasma spraying
microwave heating at 2000°C, 20 min soaking and laser sintering. Microwave sintering
time. Figure 4 shows a SEM micrograph of of W-Cu has also been successfully reported
as-pressed and microwave sintered rhenium by several researchers [Mondal et al., 2008;
compact. Mondal et al., 2009; Mondal et al., 2007].
Mondal showed that full densification can be
CONSOLIDATION OF W-Cu achieved by microwave sintering of W-15Cu
There are two ways used to form composition. In spite of higher heating rate in
W–Cu composites: copper infiltration of a microwave sintering, cracking in the sintered
porous tungsten skeleton and liquid phase samples was not observed. This is attributed
sintering (LPS) of mixed phase compacts. For to the volumetric heating nature of microwave
economical reasons, shape, complexity, and sintering. They also reported that microwave
large scale production consideration, the sintering lowers the sintering temperature
latter technique is the only widely adapted to achieve the optimum densification as
fabrication routes for W-Cu alloys. However, well as optimum mechanical properties.
the lack of solubility between tungsten and
copper makes it very difficult to achieve full
densification through liquid phase sintering.
In the LPS approach, high compaction
pressures increase the green density, but
result in increased particle-particle contacts
that limit particle rearrangement during
LPS. Other than rearrangement shrinkage,
the sintering shrinkage is due to solid state
sintering of the tungsten skeleton, even after
liquid formation. Smaller tungsten powders
induce faster skeletal sintering or allow
for a lower sintering temperature, which is
desirable since copper evaporation occurs at
high temperatures. The mutual insolubility of
the W–Cu system denies solution precipitation
controlled LPS densification. On the other
hand, the addition of small quantities of
transition elements (such as Ni, Co and Fe),
activate tungsten skeletal sintering and have
a positive impact on the sinterability of
the W–Cu system. However, these additions
degrade electrical and thermal conductivity.
Higher sintering temperature or longer
holding time always help to improve the
densification but copper may diffuse out from
the skeleton which leads to non homogeneous
microstructure. Slow heating rate, co
reduction of oxides or other thermo chemical
processes are all efficient method to improve Figure 5. SEM micrographs of (above) conventional and
the densification of this material. Novel (below) microwave sintered W-30Cu alloys, sintered at
sintering technique has also been explored to 1200°C for 30 min in H2 atmosphere.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


35
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Finer particle size and more homogeneous and heavy-duty electrical contacts. Being a
distribution of tungsten in the microstructure strategic material, all the details regarding
have been reported in microwave sintered W- the processing are not available in the open
15Cu composites. Figure 5 exhibits finer grain literature. These alloys are very structure
size and homogeneous distribution of both sensitive and therefore a firm understanding
the phases in microwave and conventionally of the processing and their physical metallurgy
sintered W-30Cu composites. are very critical for attaining better product
performance. In the past, substantial
TUNGSTEN HEAVY ALLOYS (WHAs) work has been focused on modifying the
Tungsten heavy alloy (WHA) is a group matrix composition and post-sintering heat
of two-phase composites, based on W-Ni-Cu treatment to achieve full densification and
and W-Ni-Fe. Price [Price et al., 1938] proposed optimal mechanical properties. Residual
liquid phase sintering as a viable technique porosity greater than 0.5% drastically reduces
for consolidating the W-Ni-Cu alloy. The W- the mechanical properties, especially the
Ni-Cu system provided an important basis for toughness and ductility.
understanding the phenomenology of liquid
phase sintering. In fact Cannon [Cannon and CONSOLIDATION OF W-Ni-Cu ALLOYS
Lenel, 1953] referred the mechanism of liquid In W-Ni-Cu alloys, normally the nickel-
phase sintering as ‘heavy alloy mechanism’. to-copper ratio ranges from 3:2 to 4:1. Price
WHA possesses unique combination of [Price et al, 1938] were the first to propose
properties such as high density (16-18 g/ Ni-Cu as the binder for tungsten heavy alloys.
cm3), high strength (1000-1700 MPa) and high They have shown that highly dense W-Ni-Cu
ductility (10-30%). heavy alloys could only be obtained by liquid
WHAs can be classified into two main phase sintering.
groups based on the binder composition: W- Subsequently, researchers looked
Ni-Cu and W-Ni-Fe. Besides the high density at full densified W-Ni-Cu composites by
and unique combination of high strength and solid state sintering. Solid state activated
ductility, there are other attributes, which sintering of tungsten powder with various Ni-
make WHA a versatile product [Lassner and Cu additions was studied by Brophy [Brophy
Schubert, 1999]: et al., 1966]. Their study was mainly confined
to densification mechanism. They proposed
• The high modulus of elasticity that shrinkage occurs during the initial two
• Excellent vibration damping stages of sintering: (i) rearrangement and (ii)
characteristics solution-reprecipitation stages. The role of
• Its good machinability phase relationships on the activated sintering
• The high absorption ability for of tungsten was studied by Prill [Prill, 1964].
x-rays and γ-rays They proposed that at lower Ni:Cu content,
• Good thermal and electrical the acceleration of sintering of tungsten
conductivities diminishes. Kothari [Kothari, 1967] studied
• Low electrical erosion and the densification and grain growth kinetics
welding tendency of W-Ni-Cu heavy alloys with different Ni:Cu
• Good corrosion resistance ratios.
The rate constant for the sintering
The combination of density, ductility, process was evaluated from the volume
strength, thermal conductivity and corrosion change as a function of sintering period. The
resistance makes them unique in many activation energy was found to be independent
applications such as radiation shields, of binder composition. Grain growth rate was
vibration dampers, kinetic energy penetrators proportional to the sintering period. Makarov

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


36
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

[Makarov et al., 1965] studied the coalescence employing prealloyed powders [Martin et al.,
phenomenon in W-Ni-Cu alloys during liquid 2006]. Use of prealloyed powder improves
phase sintering. The effect of tungsten and homogeneity.
copper powder size variation on the sintered The homogenization process
properties of W-Ni-Cu heavy alloys was carried accelerates sintering and promotes
out by Srikanth [Srikanth et al., 1983; Srikanth densification. The alloy formation during
and Upadhyaya, 1984]. They found enhanced sintering decreases the material viscosity and,
sintered properties with finer particle size of hence, stimulates material flow under the
the constituent with increasing nickel content action of capillary forces [Savitskii, 2005]. In
in the binder. The effect of composition and the binary Ni-W system either solid tungsten
sintering temperature on the densification or a Ni-base solid solution can co-exist in
and microstructure of W-Ni-Cu heavy alloys equilibrium with a liquid phase [Hansen and
was studied by Ramakrishanan [Ramakrishnan Anderko, 1957; Eremenko et al., 1977],
and Upadhyaya, 1990]. Kuzmic [Kuzmic, 1966] which depends upon temperature and relative
proposed that rapid cooling from the sintering amount of components in the system.
temperature prevent the formation of brittle Despite widespread application,
phase in order to obtain good mechanical difficulties still exist in the manufacture of
properties. Ariel [Ariel et al., 1973] correlated liquid phase sintered tungsten heavy alloys.
the mechanical properties of W-Ni-Cu system Large dimensional change takes place during
with sintered microstructure. Their study sintering. The linear shrinkage can be as large
showed that the mechanical properties are a as 20%, which causes slumping to occur. The
function of mean free path between tungsten distortion behavior of 80W-20 (Ni, Cu) alloy
grains, volume fraction of tungsten grains and was studied by Upadhyaya [Upadhyaya, 1998
the contiguity of tungsten spheroids. ]. He found that high dihedral angle minimize
The solubility of tungsten in the liquid compact slumping. Researchers have also
binder plays a dominant role in sintered W- attempted to consolidate W-Ni-Cu alloys using
Ni-Cu alloys. The solubility of W in the Ni-Cu microwave energy. Mondal showed that at
additive can be regulated by an appropriate relatively low temperature (1300°C) prealloyed
selection of Ni:Cu ratio. Nowadays, W-Ni- compositions gave better densification but at
Cu alloys are also being consolidated by relatively higher temperature (1450°C) which

Figure 6. SEM micrographs of (left) conventional and (rigth) microwave sintered W-7Ni-3Cu alloys. The compacts
were sintered at 1450°C [Mondal et al., 2008].

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


37
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

corresponds to the liquid phase sintering of compacts. Taking into consideration the
this system, there is no difference between slow heating rate (5°C/min) in conventional
premixed and prealloyed compositions of furnace, there was about 75% reduction in
90W-7Ni-3Cu alloys, as far as densification the process time of W-Ni-Fe compacts in
is concerned irrespective of the heating microwave furnace. Figure 7 shows a typical
mode. Figure 6 exhibits SEM micrographs of thermal profile used in their experiments. It
conventionally and microwave sintered 90W- was interesting to note that heating of the
7Ni-3Cu alloys. They also reported about 75% same compact in microwaves was achieved
reduction in process time, elimination of at much lower power consumption. This can
brittle intermetallic formation and superior be attributed to the fact that in microwaves
mechanical properties in microwave sintered the sample per se heats up and acts as the
samples. source of heat. Consequently, the effective
thermal mass reduction lowers the required
CONSOLIDATION OF W-Ni-Fe ALLOYS power input. Furthermore, for both heating
Although WHAs with Ni-Cu binder were modes, no distortion was observed in the
initially developed, attention was eventually
focussed on heavy alloys with Ni-Fe binder
because of their better mechanical properties
[25]. Conventional sintering of W-Ni-Fe alloys
results in grain growth and long sintering time
results in precipitation of brittle intermetallic
phases. To avoid this, the matrix composition
is restricted to an optimal nickel to iron ratio
of 7:3 [25]. Microwave sintering of W-Ni-Fe
composition were first studied by Upadhyaya
et al. [70]. They investigated the sintering
response of a 92.5W-7.5(Ni-Fe) alloy - with a
non-optimal matrix composition - consolidated
through microwaves and compared its
densification, microstructure and mechanical
properties vis a vis conventionally sintered

Fig. 8: SEM micrographs of (above) conventional and


Figure 7. Typical thermal profiles used for sintering (below) microwave sintered 92.5W-6.4Ni-1.1Fe alloys.
of 92.5 W -6.1Ni-1.1Fe alloys using conventional and The compacts were sintered at 1500°C [Upadhyaya et
microwave furnaces [Upadhyaya et al., 2007]. al., 2007].

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


38
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

sintered samples. Both conventional as well alloys for structural and high performance
as microwave sintered samples attained applications. Similar study on 90W-7Ni-3Fe
>98.5% density. alloys were conducted by Mondal [Mondal and
Figure 8 compares the microstructures Upadhyaya, 2007; Mondal and Upadhyaya,
of conventional and microwave sintered alloy. 2006]. They also observed the reduction in
Their observed trend is in accordance with the sintering time during the consolidation of
earlier reports of microwave advantage in this alloys. They also observed the lowering
producing finer grains, since microstructural of sintering temperature in microwave mode.
coarsening in W-Ni-Fe alloys occurs through The W-Ni-Fe alloys were sintered to equivalent
solution-reprecipitation and its kinetics level in microwaves with 75% reduction in the
is time-dependent. Hence, sintering-time sintering time.
compression in microwave furnace will Figures 10 and 11 compare the SEM
restrict W grain growth. microstructure of conventional and microwave
Figure 9 compares the XRD patterns of sintered alloys at different temperature. The
conventional and microwave sintered alloys. microwave sintered W-Ni-Fe microstructure
As expected, since the matrix was not in is less contiguous than conventional
stoichiometric proportion, therefore, phase microstructure. Microwave sintered compacts
analysis of conventionally sintered 92.5W-
6.4Ni-1.1Fe alloy shows presence of brittle
intermetallic phases (NiW, Fe7W6). Remarkably,
these phases are absent in microwave
sintered compacts. It is conjectured that
the precipitation of intermetallic phase in
conventionally sintered compacts promotes
cleavage fracture and thereby leads to
degradation of mechanical properties.
Based on the above results, it is evident
that microwave sintering offers a potentially
viable means to consolidate tungsten heavy

Figura 9. XRD analysis of microwave as well as Figure 10. SEM micrographs of (above) conventional
conventionally sintered 92.5W-6.4Ni- 1.1Fe alloy and (below) microwave sintered 90W-7Ni-3Fe alloys
[Upadhyaya et al., 2007]. sintered at 1300°C.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


39
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

have higher hardness and tensile properties


as compared to their conventionally sintered
counterparts. Figure 12 represents the XRD
analysis of conventionally and microwave
sintered samples. This analysis does not
reveal the presence of any intermetallic
phase formation irrespective of the heating
mode for this particular composition.

FINAL REMARKS
Pure refractory metals such as, W,
Re and Mo can be effectively sintered using
microwave energy to high densification.
Microwave sintering provides about 80%
reduction in total processing time. Microwave
sintering leads to higher sintered densities
(of as high as 98% of theoretical density).
Finer grain sizes and superior mechanical
properties have been achieved in microwave
sintering irrespective of the material. In case
of W sintering addition of Y2O3 and HfO2 (grain
growth inhibitors) have been successfully
used to restrict grain growth. W-based
alloys have been very effectively sintered in
microwaves without causing any micro- and
macro-cracking. In general finer particle
size and higher mechanical properties are
achieved through microwave sintering. Non-
Figure 11. SEM micrographs of (above) conventional stoichiometric compositions can be sintered
and (below) microwave sintered 90W-7Ni-3Fe alloys
using microwaves. This opens the possibility
sintered at 1450°C.
for tailoring a wide variety of compositions
for high performance applications.

ACKNOWLEDGEMENTS
This collaborative research was done
as a part of the Center for Development of
Metal-Ceramic Composites through Microwave
Processing which was funded by the Indo-US
Science and Technology Forum (IUSSTF), New
Delhi. The present project was also partially
funded by Institute of Plasma Research,
Department of Atomic Energy.

REFERENCES
Agrawal, D. (1999). “Microwave
Sintering of Ceramics, Composites, Metals,
Figure 12. XRD analysis of microwave as well as and Transparent Materials.” J Mater. Edu. vol.
conventionally sintered 90W-7Ni-3Fe alloy.
19 (4, 5, 6), pp. 49-58.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


40
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Agrawal, D., J. Cheng, P. Seegopaul, Bourell, L., and J.R. Groza


and L. Gao (2000). “Grain Growth Control (1998). “Consolidation of Ultra Fine and
in Microwave Sintering of Ultrafine WC- Nanocrystalline Powder.” in ASM Handbook
Co Composite Powder Compacts.” Powder 7 (American Society of Metals, Metals Park,
Metal., vol. 43(1), pp. 15-16. OH), pp. 583–589.
Agrawal, D.K. (1998). “Microwave Brophy, J.H. and A.L. Prill (1966).
Processing of Ceramics: A Review.” Current Opinion “A Reanalysis of Data on the Solution-
in Solid State & Mat Sci, vol. 3 (5), pp. 480-86. Reprecipitation Stage of Liquid Phase
Agrawal, D.K., A.J. Papworth, J. Sintering.” Trans. AIME, vol. 236, pp. 85-91.
Cheng, H. Jain, and D.B. Williams (2001), “ Bykov, Yu. V., K.I. Rybakov and
Microstructural Examination by TEM of WC/ V.E. Semenov (2001). “High-Temperature
Co Composites Prepared by Conventional and Microwave Processing of Materials.” J. Phys.
Microwave Processes.” Proc. 15th International D: Appl. Phys., vol. 34, pp. R55–R75.
Plansee Seminar, vol.2, G. Kneringer, P. Cannon, H.S. and F.V. Lenel (1953).
Rodhammer and P. wilhartitz (eds.),Plansee “Some Observations on the Mechanism of
AG, Reutte, Austria, pp. 677-684. Liquid Phase Sintering.” Proceedings of
Andrey, L.M., Eugene A. Olevsky Plansee Seminar, F. Benesovsky (ed.), Plansee
(2004). “Effective Diffusion Coefficients in Metallwerk, Reutte, Austria, pp. 106-122.
Solid-State Sintering.” Acta Materialia, vol. Chhillar, P., D. Agrawal and J.H. Adair
52, pp. 2953–2963. (2008). “Sintering of Molybdenum Metal
Anklekar, R.M., D.K. Agrawal, and Powder using Microwave Energy.” Powder
R. Roy (2001). “Microwave Sintering and Metallurgy , vol. 51, no. 2, pp. 182-187
Mechanical Properties of P/M Steel.” Powder Clark, D.E. and W.H. Sutton (1996).
Metal., vol. 44[4], pp. 355-362 “Microwave Processing of Materials.” Ann.
Anklekar, R.M., K. Bauer, D.K. Agrawal, Rev. Mater. Sci., vol. 26, pp. 299-331.
and R. Roy (2005). “Improved Mechanical Couchman, P.R. and W.A. Jesser (1977).
Properties and Microstructural Development “Thermodynamic Theory of Size Dependence
of Microwave Sintered Copper and Nickel of Melting Temperature in Metals,” Nature,
Steel PM Parts.” Powder Metallurgy, vol. 48, vol. 269, no.6, pp. 481-483.
no. 1, pp. 39-46. Engleman, G., J. Nable, A.J. Sherman,
Ariel, E., J. Batra, and D. Brandon L.O. Vatamanu, B. Doud and J. Stiglich (2008).
(1973). “Activated Sintering.” Powder “Development and Characterization of Nano-
Metallurgy International, vol. 5, pp 125-128. Tungsten Powder.” Proc. 2008 Intl Conf on W,
Averbach, R.S., H. Zhu, R. Tao, and H.J. Refract & Hardmetals VII, Publ MPIF, pp. 5-49
Hofler (1996). “Sintering of Nanocrystalline – 3-56.
Materials.” Experiments and Computer Eremenko, V.N., R.V. Minakova, and
Simulations, in Synthesis and Processing M.M. Churakov (1977). “Solubility of Tungsten
of Nanocrystalline Powder, edited by D.L. in Copper-Nickel Melts.” Soviet Powder
Bourell, pp. 203–216. Metallurgy and Metals Ceramics, vol. 16, pp.
Booske, J. H., R.F. Cooper, S.A. 283-286.
Freeman (1997). “Microwave Enhanced Groza, J.R. (1999). “Sintering of
Reaction Kinetics in Ceramics.” Mat. Res. Nanocrystalline Powders.” Int. J. Powder
Innovat., vol. 1, pp. 77–84. Metall. vol. 35, 59
Bose, A., B.R. Klotz, F.R. Kellogg, K.C. Groza, J.R. (2002). “Nanocrystalline
Cho and R.J. Dowding (2008). “Nanocrystalline Powder Consolidation Methods.” Nanostructured
Tungsten Powder Synthesis Using High Energy Materials-Processing, Properties and Potential
Milling.” Proc. 2008 Intl Conf on W, Refract & Applications, edited by C.C. Koch Noyes, New
Hardmetals VII, Publ MPIF, pp. 5-35 – 5-48. York, NY, pp. 115–178.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


41
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Hahn, H. (1993). “Microstructure Systems W-Ni-Fe and W-Ni-Cu.” Soviet Powder


and Properties of Nanostructured Oxides.” Metallurgy and Metal Ceramic, vol. 4, pp.
Nanostruct. Mater. vol. 2, pp. 251. 554-559.
Hansen, M., and K.P. Anderko (1957). Malewar, R., K.S. Kumar, B.S. Murty, B.
“Constitution of Binary Alloys.” McGraw-Hill, Sarma, and S.K. Pabi (2007). “On Sinterability
New York. of Nanostructured W Produced by High-Energy
Huang, H., and K. Hwang (2002). Ball Milling.” Journal of Material Research,
“Deoxidation of Molybdenum during Vacuum vol. 22, no. 5, pp.1200-1206.
Sintering.” Metall. Mater. Trans. A, vol. 33A, Martin, J.M., J.L. Johnson, R.M.
pp. 657–664. German, and F. Castro (2006). “Microstructural
Jain, M., G. Skandan, K. Martin, K. Cho, Evolution of Tungsten Heavy Alloys during
B. Klotz, R. Dowding, D. Kapoor, D. Agarwal, Heating.” San Diego.
and J. Chang (2006). “Microwave Sintering: Mishra, P., G. Sethi, and A. Upadhyaya
A New Approach to Fine-Grain Tungsten-I.” (2006). “Modeling of Microwave Heating
International Journal of Powder Metallurgy, of Particulate Metals.”Metallurgical and
vol. 42, no. 2, pp. 45-50. Materials Transactions B, vol. 37b, pp.839-
Jain, M., G. Skandan, K. Martin, K. Cho, 845.
B. Klotz, R. Dowding, D. Kapoor, D. Agarwal, Mondal, A., A. Upadhyaya and B. Sarma
and J. Chang (2006). “Microwave Sintering: (2007). “Microwave Sintering of Tungsten
A New Approach to Fine-Grain Tungsten-II.” Heavy Alloys.” International Conference on
International Journal of Powder Metallurgy, Metals and Alloys: Past, Present and Future,
vol. 42, no. 2, pp. 53-57. Kanpur, December
Johnson, J.L. (2008). “Progress Mondal, A., A. Upadhyaya and D.
in Processing Nanoscale Refractory and Agarwal (2009). “Microwave Sintering of
Hardmetal Powders.” Proc. 2008 Intl Conf on W-Cu, W-Ni-Cu and W-Ni-Fe Alloys.” 17th
W, Refract & Hardmetals VII, Publ MPIF, pp. International Plansee Seminar, 25 to 29th
5-57 – 5-71. May , Reutte, Austria
Kothari, N.C. (1967). “Densification Mondal, A., A. Upadhyaya and D.
and Grain Growth during Liquid Phase Agarwal (2009). “Sintering Advances in
Sintering of Tungsten-Nickel-Copper Alloys.” Consolidating W and its Alloys.” TMS 2009
Journal Less-Common. Metals, vol. 13, pp. Annual Meeting and Exhibition, February 15-
457-461. 19, San Francisco, California.
Kuzmic,J.F.(1966).“ModernDevelopment Mondal, A., A. Upadhyaya and D.
in Powder Metallurgy.” Ed. H.H. Hausner, Plenum Agrawal (2009). “Microwave Sintering of
Press New York, vol. 3 , pp. 166-171 W-18Cu and W-7Ni3Cu Alloys,” Journal of
Lassner, E., and W.D. Schubert (1999). Microwave Power & Electromagnetic Energy,
“Tungsten: Properties, Chemistry, Technology vol. 43, no. 1, pp. 43-1-11 – 16.
of the Elements, Alloys, and Chemical Mondal, A., and A. Upadhyaya (2006).
Compounds.” Kluwer Academic/Plenum “Sintering Advances in Consolidating W-Based
Publishers, New York Alloys.” PMRM workshop, organized by Heavy
Luo, J., C. Hunyar, l. Feher, G. Link, Alloy Penetrator Project, Tiruchirappalli.
M. Thumm, and P. Pozzo (2004). “Theory and Mondal, A., and A. Upadhyaya (2007).
Experiments of Electromagnetic Loss Mechanism “Microwave Sintering of Tungsten Heavy
for Microwave Heating of Powdered Metals.” Alloys.” PMAI Conference organized by Powder
Appl. Phys. Lett., vol.84, no.24, pp. 5076-79. Metallurgical Association of India, Delhi
Makarov, R., O.K. Teodorovich, and Mondal, A., D. Agrawal and A.
I.N. Fruntsevich (1965). “The Coalescence Upadhyaya (2008). “Effect of Microwave and
Phenomena in Liquid Phase Sintering in the Conventional Heating on Sintering Behavior of

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


42
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Tungsten Coated Copper Powder.” Proc. 2008 Prabhu, G., A. Chakraborty, B. Sarma
Intl Conf on W, Refract & Hardmetals VII, Publ (2008). “Microwave Sintering of Tungsten.”
MPIF, pp. 3-134 – 3-140. Int. Journal of Refractory Metals & Hard
Mondal, A., D. Agrawal and A. Materials,
Upadhyaya (2008). “Microwave and Price, G.H.S., C.J. Smithells, and
conventional Sintering of Premixed and S.V. Williams (1938). “Sintered Alloys. Part I.
Prealloyed Tungsten Heavy Alloys.” Materials Copper-Nickel-Tungsten Alloys Sintered with
Science and Technology, Oct 5-9, Pittsburgh, a Liquid Phase Present.” Journal of Institute.
pp. 2502-2515. Metals, vol. 62, pp. 239-264.
Mondal, A., D. Agrawal and A. Prill, A.L. (1964). “The Role of Phase
Upadhyaya (2008). “Microwave Heating of Relationships in the Activated Sintering of
Pure Copper Powder with Different Particle Tungsten.” Transaction AIME, vol. 230, pp.
Size and Porosity.” Global Congress on 769-772.
Microwave Energy Application, Japan, pp. Ramakrishnan, K.N., and G.S.
517-520 Upadhyaya (1990). “Effect of Composition
Mondal, A., D. Agrawal and A. and Sintering on the Densification and
Upadhyaya (2008). “Microwave Sintering of Microstructure of Heavy Alloys Containing
Tungsten Based Alloys.” Proc. 2008 Intl Conf Copper and Nickel.” Journal of Materials
on W, Refract & Hardmetals VII, Publ MPIF, Science Letters, vol. 9, pp. 456-459.
pp. 3-122 – 3-132. Rao, K.J. and P.D. Ramesh (1995).
Mondal, A., D. Agrawal and A. “Use of Microwaves for the Synthesis and
Upadhyaya (2008). “Sintering Advances Processing of Materials.” Bull. Mater. Sci.,
in Consolidating W Based Alloys.” Global vol.18, no.4, pp. 447-465.
Congress on Microwave Energy Application, Rodiger, K., K. Dreyer, T. Gerdes, and
Japan, pp. 301-304. M.W. Porada (1998). “Microwave Sintering
Mondal, A., D. Agrawal and A. of Hardmetals.” International Journal of
Upadhyaya (2009). “Microwave Heating of Refractory Metals & Hard Materials, vol.16,
Pure Copper Powder with Varying Particle Size pp. 409-416.
and Porosity.” Journal of Microwave Power & Roy, R., D.K. Agrawal, J.P. Cheng, and
Electromagnetic Energy, vol 43, no. 1, pp. 43- S. Gedevanishvili (1999). “Full Sintering of
1-5 – 10. Powdered Metals Using Microwaves.” Nature,
Mondal, A., D. Agrawal, A. Upadhyaya, vol.399, no.17, pp. 668-670.
P. Chhillar, J. Cheng, and R. Roy (2009). Rybakov, K.I., V.E. Semenov, S.V.
“Microwave Sintering of Refractory Metals: W, Egorov, A.G. Eremeev, I.V. Plotnikov, and
Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe Alloys.” Y.V. Bykov (2006). “Microwave Heating of
International Conference on Materials for Conductive Powder Materials.” J. Appl. Phys.,
Advanced Technology, 28th June to 3rd July, vol. 99, pp. 023506-09.
Singapore. Saitou, K. (2006). “Microwave Sintering
Oda, E., K. Ameyama, S. Yamaguchi of Iron, Cobalt, Nickel, Copper and Stainless
(2006). “Fabrication of Nano Grain Tungsten Steel Powders.” Scripta Materialia, vol. 54,
Compact by Mechanical Milling Process and no. 5, pp. 875-879.
its High Temperature Properties.” Materials Savitskii, A.P (2005). “Scientific
Science Forum, vol. 503-504, pp 573-578. Approaches to Problems of Mixtures Sintering.”
Patrician, T., V. Sylvester and R. Daga Science of Sintering, vol. 37, pp. 03-17.
(1985). Proc. Symp. On “Physical Metallurgy Srikanth, V. and G.S. Upadhyaya
and Technology of Molybdenum and its Alloys”; (1983). “Properties of Sintered Heavy Alloys.”
Ann Arbor, MI, USA, AMAX Materials Research International Journal of Refractory and Hard
Center, 1–11. Metals, vol. 2, no 3. pp. 49-54.

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


43
International Microwave Power Institute
Avijit Mondal et al., Microwave Sintering of Refractory Metals/alloys: W, Mo, Re, W-Cu, W-Ni-Cu and W-Ni-Fe ...

Srikanth, V. and G.S. Upadhyaya Upadhyaya, A. (2001). “Processing


(1984). “Effect of Tungsten Particle Size on Strategy for Consolidating Tungsten Heavy
Sintered Properties of Heavy Alloys.” Powder Alloys for Ordnance Applications.” Materials
Technology, vol. 39, pp. 61-65. Chemistry and Physics, vol. 67, pp. 101-110.
Sutton, W. (1992). “Microwave Upadhyaya, A., (1998). “A
Processing of Ceramics-an Overview.” Mater Microstructure-Based Model for Shape
Res Sot Symp Proc, vol. 269, pp. 3-l 9. Distortion during Liquid Phase Sintering.”
Takayama, S., G. Link, S. Miksch, M. Ph.D.Thesis, Penn State, USA.
Sato, J. Ichikawa, and M. Thumm (2006). Upadhyaya, A., S. K. Tiwari, and P.
“Millimetre Wave Effects on Sintering Mishra (2007). “Microwave Sintering of W-Ni-
Behavior of Metal Powder Compacts.” Powder Fe Alloys.” Scripta Materialia, vol.56, pp. 5-8.
Metallurgy, vol.49, no.3, pp. 274-280. Von Hippel, A. (1954). “Dielectric
Tinga, W.R. and A.G. Voss (1968). Materials and Applications,” Technology Press
“Microwave Power Engineering.” New York: of M.I.T, p.p.301
Academic Wang, H., Z.Z. Fang and D. Siddle
Tinga, W.R. and E.M Edwards (1968). (2008). “Study of Size-Dependent Sintering
“Dielectric Measurements Using Swept Behavior of Tungsten Powders.” Proc. 2008
Frequency Techniques.” Journal of Microwave Intl Conf on W, Refract & Hardmetals VII, Publ
Power, vol.3, no. 3, p.p.114-125. MPIF, pp. 5-72 – 5-77

Journal of Microwave Power and Electromagnetic Energy, 44 (1), 2010


44
International Microwave Power Institute
Science of Sintering, 42 (2010) 169-182
________________________________________________________________________

doi: 10.2298/SOS1002169M

UDK 622.785:546.56
Effect of Porosity and Particle Size on Microwave Heating
of Copper
A. Mondal1, A. Shukla1, A. Upadhyaya1, *), D. Agrawal2
1
Department of Materials & Metallurgical Engineering Indian Institute of
Technology, Kanpur 208016, India
2
Materials Research Institute, The Pennsylvania State University, University Park,
PA 16802, USA

Abstract:
The present study investigates the effect of varying particle size and porosity on the
heating behavior of a metallic particulate compact in a 2.45GHz multimode microwave
furnace. Experiments on copper suggest that unlike monolithic (bulk) materials, metallic
materials do couple with microwaves when they are in particulate form. The powder
compacts having higher porosity and smaller particle sizes interact more effectively with
microwaves and are heated more rapidly. A dynamic electromagnetic-thermal model was
developed to simulate the temporal temperature distribution using a 2-D finite difference time
domain (FDTD) approach. The model predicts the variation in temperature with time during
heating of copper powder compacts. The simulated heating profiles correlate well with those
observed from experiments.
Keywords: Finite Difference Modelling (FDM); Microwave heating; Sintering; Numerical
Modeling, Heat Transfer

Introduction

Microwaves are primarily coherent, polarized electromagnetic waves which lie


between the radio-waves and visible light in the electromagnetic spectrum. Their frequency
lies between 0.3 to 300 GHz and wavelength varies from 1 mm to 1 m range. One of the key
applications of microwaves is to heat materials. For such applications, however, only few
frequencies are allowed, namely 915 MHz, 2.45 GHz, 5.8 GHz and 28 GHz [1]. Out of these
2.45 GHz multimode microwaves are most commonly used. For engineering applications,
microwaves are known to interact strongly with a range of ceramics – both in bulk as well as
particulate form - and heat them to high temperatures (up to 1800°C) [2]. Unlike conventional
heating, wherein the heat-transfer is limited by radiative and conductive mode, microwaves
directly couple with ceramics result in rapid and uniform heating. Microwave heating of
ceramic and other dielectrically lossy materials have been widely investigated and the
mechanism of microwave-material interaction is well documented [2-4]. Microwave
penetrates and propagates through dielectric material. Thus, in effect, the material per se
becomes the source of heat. Consequently, at times, the flow of heat is the reverse of that
observed in conventional heating.
Unlike ceramics, bulk or monolithic metallic materials are known to reflect
_____________________________
*)
Corresponding author: anishu@iitk.ac.in
170 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

microwaves at room temperature. Being good conductors, no internal electrical field is


induced in metallic materials. The induced electrical charges remain restricted at the surface
of a bulk metallic sample. This causes the eddy currents and subsequently resistive losses.
Roy et al. [5] were the first to recognize that this can be exploited to heat metals too to high
temperatures provided they are in particulate form. This opened the possibility of
consolidating the particulate metal compacts to full density through sintering using
microwaves. Subsequently, several researchers [6-19] have demonstrated that a range of
particulate metals and alloys can be heated using microwaves and can result an overall 60 to
90% reduction in the processing time. Besides cost advantage, a reduced processing time
restricts microstructural coarsening that occurs concomitantly with densification during
sintering. Hence, as compared to compacts processed in a conventional (radiatively-heated)
furnace, microwave sintered compacts have more homogenous and refined microstructure
which leads to an improved mechanical properties [8-19].
While the efficacy of microwaves to sinter metals is now well recognized, a
systematic evaluation of factors influencing the microwave interaction with the particulate
compact and its influence on the heating rate is as yet lacking. Recently, Saitou [10]
systematically investigated microwave heating of several metal powder compacts and
reported that the sintering mechanism per se remain the same both in conventional as well as
microwave heating. In recent years, using single-mode microwaves where the electric (E) and
magnetic (H) fields were separated; researchers [20-23] have demonstrated experimentally
that the effect of microwave frequency, sample conductivity, and green density on heating
efficiency of metallic compacts. However, from an industrial materials processing standpoint,
single-mode microwaves have limited application. Mishra et al. [24] modeled microwave
heating in metallic systems by solving the Maxwell’s equation and using a two-dimensional
finite difference time domain (FDTD) technique. However, their model ignored the influence
of the initial compact porosity and was not validated experimentally.
From the preceding discussion, it is evident that while microwave heating of ceramics
is well-understood and many models have been attempted, there is still an uncertainty about
the exact mechanism and mode of microwave heating in case of particulate metals. This study
therefore sets out to systematically investigate - both experimentally as well as through a
thermal-electromagnetic model - the effect of initial powder size and the as-pressed porosity
on the heating response of a copper compact in a multimode (2.45 GHz) microwave furnace.

2. Modeling approach

Modeling of microwave heating involves solving Maxwell equations of


electromagnetism simultaneously with the heat transfer equation [25]. The detailed approach
which has been adapted in this study to calculate the electro-magnetic absorption (PEM) by
metal powder compact has been discussed elsewhere [24].
In the present study, the calculated power absorbed is correlated with the temperature
rise in the compact. The latter requires solving the heat transfer equation for the boundary
conditions defined by the dimensions of the sample and its physical, thermal and dielectric
properties. The dependency of these parameters on temperature necessitates the need for a
dynamic simulation. Figure 1 shows the modeling approach adopted in this study through a
schematic flow-chart.
Microwaves interaction with metals is restricted to its surface only. This depth of
penetration in metals, also known as skin-depth (δ), is defined as the distance into the material
at which the incident power drops to 1/e (36.8%) of the surface value. The skin depth is
mathematically expressed as follows [26]:

δ = 1
= 0 . 029 ρλ o (1)
π f μσ
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 171
___________________________________________________________________________

where, ƒ is normal microwave frequency, μ is magnetic permeability, σ is electrical


conductivity, ρ is electrical resistivity, λo is incident wavelength (12.24 cm for 2.45 GHz
waves). The skin-depth in metals typically varies between 0.1 to 10 μm. From Equation 1, it
is evident that metals with higher electrical conductivity have lower skin depths. For metals,
as the resistivity increases with increase in temperature, the skin depth too increases. In
general, the skin depth is relatively small in metals; however, most metallic powders typically
have equivalent dimensions. In metal powders, the surface area and thereby the ‘effective
skin’ (portion of metal powder that couples with microwaves) is high enough to contribute to
its heating. Thus, it is likely that some of the submicron and nano-sized metal powders
undergo volumetric heating when subjected to microwaves. The skin depth of the metal plays
an important role in governing the power loss during the microwave-metal interaction which
leads to its heating [25]. The tangential component of the magnetic field, Ht, of microwaves
induces an electrical field, E, at the metal powder surface. The induced electrical field
generates surface current which causes the resistive Joule heating in metal powders.
Assuming spherical metal powder of radius, rp, the electro-magnetic power density (PEM) can
be expressed as [25]:
2
6 Rs Ε o
Ρ EM = 3Ρ
rp = η o 2 ⋅r p
(2)

where, Rs is surface resistivity ( Rs = 1/σδ, σ is electrical conductivity and δ is skin depth); Eo


is electric field amplitude at surface; and ηo is impedance of free space (377 Ω). In our
approach Eo is considered as a tunning parameter to fit the experimental data. In our earlier
work [24] the effect of porosity and particle size has not been considered whereas in this
model it has been considered.

2.2. Assumptions

To make the problem tractable and simplistic for modeling, following assumptions
are made:
• Since the particle size is much smaller than the wavelength of the microwave
radiation, the field across the particle is assumed to be spatially uniform and periodic in time.
• Powder coarsening is not assumed during heating.
• Monosized powders are assumed so that the calculated EM-power density generated
in a powder will be same when integrated over bulk aggregate.
• All metal powders within a compact are assumed to heat-up at the same rate so no
heat conduction between particles occurs.
• A heat-factor is considered to account for the presence of susceptor.
• Effect of oxide layer formation is not considered in the present model.
As shown schematically through Fig. 1, Eo which is the electric field amplitude at surface are
in turn fed as input parameter into the thermal model.

2.3. Thermal Model for Predicting Temperature Rise

Thermal Model for Cylindrical Geometry Compact has been defined as following:

∂T 1 ∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞
ρC P = ⎜ Kr ⎟+ ⎜K ⎟ + PEM (3)
∂t r ∂r ⎝ ∂r ⎠ ∂z ⎝ ∂z ⎠

where
172 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

ρ : Density of the material, which is time dependent and increases as densification


increases.
K: Thermal conductivity of the material (temperature dependent)
C P : Specific heat of the material (temperature dependent)
PEM : Power density due to microwave effect (temperature dependent)

Fig. 1. Flow chart of the dynamic model proposed for simulating the thermal profile in the
present study.

Typically, in the initial stage of heating, the metal powder compacts require
susceptor-assisted heating [27]. The susceptors usually couple very well with the microwaves
and are used for initially raising the temperature of the compact. Beyond a critical
temperature, the heating occurs primarily due to microwave-metal powder surface interaction.
Besides, assisting heating in the initial stage, the main role of susceptor is to prevent heat loss
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 173
___________________________________________________________________________

from the surface of the powder compact to the surrounding through radiation. In the present
modeling scheme, the contribution of SiC as susceptor for compact heating was also included.
Hence, the boundary condition at the surfaces of the compact can be formulated as following:

∂T
= h(T − Ta ) + K SUC (T − TSUC ) + σεF (T 4 − Tencl )
4
−K (4)
∂r

where
H: Heat Transfer Coefficient for convective heat transfer to surrounding
TSUC : Temperature of the SiC suspector
Tencl : Temperature of the enclosure in which compact is put
Ta : Ambiant Air temperature
K SUC : Susceptor constant
σ : Stefens-Boltzman’s constant
ε : Emissivity
F : View Factor
When no thermal gradient inside the compact exists, above equations may be written as
following:

⎛ V ⎞ ∂T ⎛V ⎞
− ρC P ⎜ ⎟ = h(T − Ta ) + K SUC (T 4 − T 4 SUC ) + σεF (T 4 − Tencl ) − PEM ⎜ ⎟
4
(5)
⎝ A ⎠ ∂t ⎝ A⎠

where
V πR 2 H
=
(
A 2πRH + 2πR 2 )
h is convective heat transfer coefficient of incoming gas,
A/V is surface area to volume ratio for compact,
σ is Stefan Boltzman constant,
ε is effective emissivity of metal powder, and
Ta is surrounding temperature.
Above equation may be discretisized with respected to time in following manner:
Tn +1 − Tn ∂T
= (6)
Δt ∂t
In Equation 6, Δt should not be too small as it increases the number of iterations, hence, the
round-off error is also increased. However, too large a Δt should also be avoided as it causes
an increase in the truncation error per step. Therefore, a suitable value of Δt is chosen which
satisfies stability criterion given below:
0.05 ≤ ∆t٠K ≤ 0.1 (7)
where, K is upper bound of following expression:
⎧ A ⎫ 1
d ⎨ ρC p ( (h(T − Ta ) + K SUC (T − TSUC ) + σεF (T 4 − Tencl
4
) − PEM )⎬
⎩ V ⎭ dT

3. Experimental setup to validate model

The as-received gas-atomized copper powders (supplier: American Chemet Corp.


USA) of different particle size were characterized for their size, size distribution and
174 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

morphology. Tab. I summarizes the characteristics of the as-received powders used for this
study. Fig. 2 shows the scanning electron micrographs of the powders in as-received
condition. The particle size analysis of the as-received powders was performed using a
Malvern Instruments particle size analyzer which works on the principle of laser diffraction.

Tab. I. Characteristics of the as-received Cu powders used for the present study.

Cu Powder Grade A B C D E
Purity, % ~99 >99.5 >99.2 >99.2 >98
Particle size, µm
D10 2 6 8 26 231
D50 6 12 18 63 383
D90 36 20 33 118 560
Specific surface area, 2.5 1.4 0.5 0.3 0.3
2
/

Fig. 2. SEM micrograph Cu powders used in the present study and having average particle
size of (a) 6 µm, (b) 12 µm, (c) 18 µm, (d) 63 µm and (e) 383 µm, respectively.

The surface areas of the as received powders were measured using a BET apparatus.
The morphological analyses of the powders were carried out using SEM imaging. Tab. II
summarizes the physical and thermal properties of the copper powder used for the present
investigations, which are the input parameters in the model. The total experimental approach
consists of two sub sets. First experiment is all about the effect of particle size during the
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 175
___________________________________________________________________________

heating of powder particle in a multimode microwave furnace. Second experiment includes


the role of initial density/porosity during the heating of metal powders in a multimode
microwave furnace. For the first experiment powders were uniaxially pressed at 400 Mpa in
to compacts with green densities ranging between 60 to 83%Th. The sintering response on
densification and microstructures were evaluated on cylindrical pellets (12.7 mm diameter
and 5 mm average height). For the second set of experiments cylindrical pellets of the same
dimensions of as-received copper powder (particle size 18 µm) were compacted using the
varying compaction pressure in the range of 90 to 350 Mpa to obtain green densities ranging
between 56 to 76%Th.

Tab. II. Physical and thermal properties of Cu used for simulation of the thermal profile
[28,29].
Density, gcm-3 8.9
Heat capacity (Cp), Jkg-1K-1 92•10-3T + 355.9
Electrical resistivity, μΩ-cm 6.9×10-3 (T-273) + 1.5
Emissivity 0.65

Fig. 3. Schematic diagram showing the inside of the multimode microwave furnace used in
the present study.

Microwave heating of the as-pressed compacts was carried out using 2.45 GHz
multimode microwaves in a 6kW capacity furnace. The experimental setup for the furnace is
shown in Figure 3. A multi- layered insulation package was used to provide sufficient
insulation to obtain high and uniform temperatures throughout the sample. The outer package
was made up of thick ceramic fiber (aluminum silicate) sheets. A mullite tube was placed at
the centre of the package, and samples were placed inside this mullite tube. The entire
package was placed on a turntable to ensure uniform exposure of the sample to the microwave
field. Further details about the experimental setup are described elsewhere [27]. The
temperature of the sample was monitored using an infrared pyrometer (Type: RAYMA25CCF
Manufacturer: Raytek Co., Santa Cruz, CA, USA). Temperature measurement through
pyrometer is emissivity-based. Typically, emissivity varies with temperature. However, as
176 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

very little variation in the emissivity was reported in the temperature range used in the present
study, hence, the effect of variation in emissivity was ignored in the present investigation.
The sintered density was obtained by both dimensional measurements as well as
Archimedes density measurement technique. The sinterability of the compact was also
determined through densification parameter which is expressed as:
(sintered density - green density)
Densification parameter = ……… (8)
(theoretical density - green density)

4. Results and discussion

In the preceding sections, a theoretical approach has been formulated to predict the
heating profile of metal powder compacts in microwave furnace. One of the first objectives
was to test the efficacy of the model to predict the temperature distribution within the
compact during heating. To test this, thermal profiles was simulated for Cu powder compact
(particle size 18 µm) having radius 6.5 mm and 6.5 cm respectively (Fig. 4).

Fig. 4. Simulated isothermal contours as a function of time in cylindrical Cu compacts(Avg.


particle size 18 µm) having dimensions (a) 6.5 mm radius and 5.5 mm height; and (b) 6.5 cm
radius and 5.5 cm height, respectively.
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 177
___________________________________________________________________________

As is evident from Figs. 4a and 4b, for copper powder compacts, irrespective of the compact
size, the temperature remains homogenous throughout its cross-section. It is also interesting to
note that for both cases, the temperature homogenization is achieved within a few seconds.
Therefore the assumption that the temperature is uniform inside the compact is valid.
1200

1000
Temperature, (°C)

800

600
6 µm
12 µm
400 18 µm
63 µm
383 µm
200 bulk Cu

0
0 10 20 30 40 50 60 70
Time, (min)

Fig. 5. Effect of varying particle size on the heating behavior of Cu compacts in a microwave
furnace. For comparison the thermal profile of bulk Cu under microwave is also
superimposed.

1200

1000
Temperature, (°C)

800

600 % Porosity
44
400 35
29
24
200

0
0 2 4 6 8 10 12 14
Time, (min)

Fig. 6. Effect of varying initial porosity on the heating behavior of Cu compacts (avg. size: 18
µm) in a microwave furnace.

Figs. 5 and 6 compare the effect of the initial powder size and the as-pressed porosity
on the thermal profiles of cylindrical copper compacts exposed to microwaves. For the sake
of comparison, the thermal profile on bulk copper is also superimposed in Figure 5. As
expected, the bulk copper does not heat up at all in microwave. In contrast, particulate
compacts show strong interaction with microwaves. It is evident from the figures that the
microwave heating is more effective for compacts having smaller particle size and high
178 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

porosity levels. Elsewhere, Ma et al. [23] too have reported similar observations in copper
powder compacts heated in single-mode microwaves.
As particle size increases the heating rate decreases and after certain time for a
particular particle size; heating rate becomes constant at a particular power setting. It is well
known that microwaves for electrically conducting materials such as copper do not penetrate
a bulk sample beyond the skin depth. The skin depth expression in Equation 1 can be
rewritten as:
2
δ= (9)
μσω
where, µ is the real part of the permeability taken to be 4 ×10-7 T m/A for nonmagnetic
materials. σ and ω are the dc electrical conductivity and angular frequency, respectively. At
2.45 GHz, the skin depth of bulk copper is about 1.3 µm. However, in case of porous
compacts, the electrical conductivity correspondingly decreases. Thus, from the above
equation effective skin depth also increases proportionately. The variation in the heating rates
therefore can be attributed to the difference in skin depth and also the change in surface area
per unit volume with the varying particle sizes. This effect is further manifested to a larger
extent for coarser particle sizes on account of their higher compressibility. This is confirmed
by recently report from Ma et al. [23] who have experimentally shown that the electrical
conductivity of the as-pressed copper powder compacts with 22 µm powder is 104 times
higher than those prepared using 3 µm powder.

Tab. III: Densification response of Cu powder of different particle under microwave heating.

Average Cu Powder Size (µm) 6 12 18 63 383

sintered density (% theoretical) 95.4 90.4 90.0 89.8 90.0


densification parameter 0.88 0.54 0.48 0.43 0.39

Tab. IV. Effect of compaction pressure on the densification response of Cu powder compact
having average particle size of 18 µm.
Compaction Pressure (MPa) 90 210 280 350
green density (% theoretical) 56 65 71 76

sintered density (% theoretical) 78 82 87 90

densification parameter 0.50 0.49 0.55 0.58

Tab. III compares the densification response of all the sintered copper compacts of
different particle sizes in microwave furnace. In case of smallest particle compacts,
densification during microwave sintering is highest and the trend is similar for the rest of
other particle sizes. This is also validated by comparing the densification parameters which
follows similar trend as sintered density. Tab. IV compares the densification response of
different initial green density copper compacts sintered in microwave furnace. Although the
heating rate is faster for lower green density samples than the higher green density compacts
but densification during microwave sintering is higher in case of higher green density
compacts and the densification trend is also similar. This can be attributed to the difference in
the initial density.
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 179
___________________________________________________________________________

Fig. 7. Effect of varying initial particle size [(a) 6 µm, (b) 12 µm, (c) 18 µm, (d) 63 µm, (e)
383 µm] on the actual and predicted thermal profiles (left) and their correlation assuming a
linear fit (right).

Figs. 7a to 7e compare the actual and predicted thermal profile of copper powder
compact with increasing powder size. For all the five powder sizes, the trends in experimental
and predicted heating profile are self-similar. The differences in experimental and simulated
heating profiles can partly be attributed to the fact that power input was kept constant in
modelling, whereas, in the actual experiment, it was changed from 0.5 kW to 1.3 kW during
this experiment. An interesting observation for both experimental and simulation of thermal
profile is that for a particular choice of experimental variables, the temperature rise is
180 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

restricted to a certain level. This can be attributed to the fact that the heat generated in the
compact (due to electromagnetic power absorbed) is balanced by the convective and radiative
heat losses. The figures to the right are the corresponding regression fit for the actual and
predicted temperatures for each time interval. The figures and the corresponding coefficient
of correlation (R2) values suggest that for lower particle sizes the model prediction are closer
to those measured experimentally because smaller particles which size is comparable with the
skin depth undergoes volumetric heating and much better microwave effect whereas powders
with larger particle size undergoes only surface heating and lesser microwave induced heating
effect. Smaller particle attains higher maximum temperature as compared to the larger
particle.

Fig. 8. Effect of varying initial porosity in a Cu compact [(a) 44%, (b) 35%, (c) 29%, (d)
24%] on the actual and predicted thermal profiles (left) and their correlation assuming a linear
fit (right).
A. Mondal et al./Science of Sintering, 42 (2010) 169-182 181
___________________________________________________________________________

Figs. 8a to 8d compare the experimental and simulated thermal profiles of copper


powder compacts (particle size: 18 µm) having initial porosity of 44%, 35%, 29% and 24%,
respectively. The figures to the right are the corresponding predicted and actual temperature
correlation plots. It is worth noticing that for varying porosity levels the model predictions are
quite close to the experimental observations. For all the cases the correlation coefficient is
greater than 0.95.
Unlike microwave heating, for conventional furnace sintering, high green density
compacts are preferred so as to ensure homogenous heat transfer. The high green density
requires larger compaction pressure which results in lower tool life. The relatively longer
sintering time during conventional sintering is accompanied by microstructural coarsening
that is detrimental to the mechanical properties. For small sized particle compacts, the
problem is further compounded during conventional sintering since they have poor
compressibility and are also more prone to grain coarsening. From the above discussion, it is
logical to conclude that a finer powder size and higher as-pressed porosity compacts will
respond more favourably when subjected to microwave heating. The present study therefore
provides guidelines for effective design and optimization of process parameters to attain
greater microwave heating efficacy.

5. Conclusions

This study shows that copper powder compacts strongly couple with micowaves and
can be rapidly heated to high temperatures. The heating rate of the particulate compacts
increases as the particle size and green density decrease. A two-dimensional FDTD approach
was employed to simulate the heating response of the powder compacts that incorporated both
the electromagnetic and thermal model. A special feature of this approach is that it considers
the temperature-dependency on the physical and thermal properties. The predicted thermal
profiles correlate well with the experimental observations.

6. Acknowledgements

This work has been partially supported by Indo-US Joint Centre on Microwave
Applications funded under the program of Indo-US Public-Private Networked Joint Center
Program, India, and National Institute for Fusions Science, Japan.

7. Reference

1. A.C. Metaxas: Foundation of Electroheat: A Unified Approach, J. Wiley & Sons


Inc., New York, NY, 1996.
2. W. Sutton,: Mater. Res. Soc. Symp Proc. 269 (1992) 3.
3. D.E. Clark, W.H. Sutton, Ann. Rev. Mater. Sci. 26 (1996) 299.
4. D.K. Agrawal, Current Opinion in Solid State Mater. Sci. 3 (1998) 480.
5. R. Roy, D.K. Agrawal, J.P. Cheng, S. Gedevanishvili, Nature 399 (1999) 668.
6. D. Agrawal, J. Cheng, P. Seegopaul, L. Gao, Powder Metall. 43 (2000) 15.
7. K. Rödiger, K. Dreyer, T. Gerdes, M.W. Porada, Int. J. Ref. Metals Hard Mater.
16 (1998) 409.
8. R.M. Anklekar, D.K. Agrawal, R. Roy, Powder Metall, 44 (2000) 355.
9. G. Sethi, A. Upadhyaya, D. Agrawal, Sci. Sintering 35 (2003) 49.
10. K. Saitou, Scripta Mater. 54 (2006) 875.
182 A. Mondal et al. /Science of Sintering, 42 (2010) 169-182
___________________________________________________________________________

11. S. Takayama, G. Link, S. Miksch, M. Sato, J. Ichikawa, M. Thumm, Powder


Metall. 49 (2006) 274.
12. A. Upadhyaya, S.K. Tiwari, P. Mishra, Scripta Mater. 56 (2007) 5.
13. W.L. Wong, M.Gupta, Composites Sci. Tech. 67 (2007) 1541.
14. P. Chhillar, D. Agrawal, J.H. Adair, Powder Metall. 51 (2008) 182.
15. A. Mondal, D. Agrawal, A. Upadhyaya, in Proc. 2008 Int. Conf. Tungsten,
Refractory Hardmetals VII, Metal Powder Industries Federation, Princeton, NJ, v.
3, pp. 122-132.
16. A. Mondal, D. Agrawal, A. Upadhyaya, Mater. Sci. Tech. 2008, 2502.
17. A. Mondal, A. Upadhyaya, D. Agrawal, J. Microwave Power Electromag. Energy
43 (2009) 11.
18. A. Mondal, D. Agrawal, A. Upadhyaya, J. Microwave Power Electromag.
Energy, 44 (2010) (in press).
19. G. Prabhu, A. Chakraborty, B. Sarma, Int. J. Ref. Metals Hard Mater. 27 (2009)
545.
20. J. Cheng, R. Roy, D. Agrawal, Mat. Res. Innovations 5 (2002) 170.
21. J. Luo, C. Hunyar, I. Feher, G. Link, M. Thumm, P. Pozzo, Appl. Phy. Lett. 84
(2004) 5076.
22. K.I. Rybakov, V.E. Semenov, S.V. Egorov, A.G. Eremeev, I.V. Plotnikov, Y.V.
Bykov, J. Appl. Phy.99 (2006) 023506.
23. J. Ma, J.F. Diehl, E.J. Johnson, K.R. Martin, N.M. Miskovsky, C.T. Smith, G.J.
Weisel, B.L. Weiss, D.T. Zimmerman, J. Appl. Phy. 101 (2007) 074906.
24. P. Mishra, G. Sethi, A. Upadhyaya, Metall. Mater. Trans. B 37B (2006) 839.
25. J.H. Hill, T.R. Merchant, Appl. Mathematical Modelling 20 (1996) 3.
26. M. Pozar, Microwave Engineering, 2nd ed., J. Wiley & Sons Inc., Toronto,
Canada, 2001.
27. A. Mondal, D. Agrawal, A. Upadhyaya, J. Microwave Power & Electromag.
Energy 43 (2009) 5.
28. E.A. Brandes, and G.B. Brook: Smithells Metals Reference Book, 7th ed.,
Butterworth Heinemannn, Oxford, UK. 2000.
29. A. Cezairliyan, J.L. McClure, J. Res. National Bureau of Standards 75A (1971)
283.

Садржај: У овом раду проучен је утицај варирања величине честице и порозности на


понашање компакта металног партикулата током загревања у 2.45 GHz
мултимодалној микроталасној пећи. Експерименти вођени на бакру предлажу да за
разлику од монолитичних (балк) материјала, метални материјали се куплују са
микроталасима када су у форми партикулата. Компакти праха са већом порозношћу и
мањим величинама ћестица реагују ефективније са микроталасима и загревају се
брже. развијен је динамички електромагнетни-термички модел за симулирање
темпоралне расподеле температуре коришћењем приступа дводимензионалних
коначним разликама временских домена. Модел предвижа варијацију температуре са
временом током загревања компакта од бакарног праха. Симулирани профили
загревања су добро корелисани са експерименталним подацима.
Кључне речи: Моделирање коначним разликама, микроталасно загревање,
синтеровање, нумеричко моделовање, пренос топлоте.
MICROWAVE HEATING OF PURE COPPER POWDER
WITH VARYING PARTICLE SIZE AND POROSITY

Avijit Mondal1*, Dinesh Agrawal2 and Anish Upadhyaya1


Department of Materials & Metallurgical Engineering
1

Indian Institute of Technology, Kanpur 208016, India


2
The Pennsylvania State University, University Park, PA 16802, USA
*
avijitm@iitk.ac.in

In recent years, microwave processing of metal/alloy powders have gained considerable


potential in the field of material synthesis. Microwave heating is recognized for its various advan-
tages such as: time and energy saving, rapid heating rates, considerably reduced processing cycle
time and temperature, fine microstructures and improved mechanical properties, better product
performance, etc. Microwave material interaction for materials having bound charge are well es-
tablished, but for highly conductive materials like metals, there is not much information available to
interpret the mechanism of microwave heating and subsequent sintering of metallic materials. The
present study describes how the thermal profile of electrically conductive powder metal like copper
changes with particle size and also with porosity content; in other words, initial green density when
the material is exposed to 2.45 GHz microwave radiation in a multimode microwave furnace.

Submission Date: 5 August 2008


Acceptance Date: 29 December 2008
Publication Date: 16 January 2009

INTRODUCTION graded materials, metals and alloys [Rodiger et


al., 1998; Porada and Borchert, 1996-97; Jain
Microwave processing technology has been et al., 2006; Sethi et al., 2003; Anklekar et al.,
developed in the field of ceramics and compos- 2001].
ites. Microwave heating allows an instantaneous Microwave sintering of metal powders
volumetric, and therefore a more rapid, heating which have high electrical conductivity is a
in comparison with the conventional heating new area with growing interest. It was first
process. It also provides enhanced sintering reported in 1999 by Roy and coworkers [1999]
kinetics. As a result, it allows a reduction of that a porous, powder metal compact could be
process temperature and time which leads to heated and sintered in a microwave field. At that
an increased productivity and a reduction of time, this was considered surprising because the
energy consumption. During the last decade, the electrically conducting materials were supposed
microwave has been used for sintering particu- to reflect microwave radiation. Later, other
late metals. The reported materials of interest researchers also demonstrated that all powder
included hard metals, metal–ceramic functional metals at room temperature absorb microwaves
and only bulk metals reflect the microwaves,
Keywords: Cu metal powder, microwave sintering, allowing only surface penetration.
particle size, porosity
However, these empirical studies did not
Guest Editor: Dr. Satoshi Horikoshi, Tokyo University of attempt to explain the observed heating trends
Science, Chiba, Japan and presently there are few reports regarding

International Microwave Power Institute 43-1-5


the details of the direct interaction of micro- an increase in skin depth.
waves with powder metal compacts. Sintering Many researchers have addressed the role of
mechanisms and activation energy for diffusion several parameters such as sample conductivity
were studied by Saitou [2006] and he concluded and frequency [Ma et al., 2007], and effect of E
that microwave radiation does not change the and H field [Rybakov et al., 2006] upon thermal
sintering mechanism for diffusion in the case profile which affect the microwave absorption
of metallic material sintering. Thus, there is not by pure powdered metals. In the literature, the
much information available by which to inter- researchers have also addressed the role of par-
pret the mechanism of microwave heating and ticle size and initial porosity upon the thermal
subsequent sintering of metallic materials. profile during microwave heating of different
Microwave heating of ceramic and other conductive metal powders [Rybakov et al., 2006;
dielectrically lossy materials have been widely Mishra et al., 2006; Vaidhyanathan and Rao,
investigated and the mechanism of micro- 1997].
wave-material interaction is well documented. This paper reports how the thermal profile
Microwave penetrates and propagates through of electrically conductive metal like copper
dielectric material, such as SiC. This generates changes with particle size and also with poros-
an internal electric field (E) within a specific ity content (i.e., initial green density) when the
volume, which in turn induces polarization and material is exposed to 2.45 GHz microwave
movement of charges. The resistance to these radiation in a multimode microwave furnace.
induced motions due to internal, electric and
frictional forces attenuates the electric field. EXPERIMENTAL PROCEDURE
These losses result in volumetric heating.
The microwave-metal interaction is quite For the present study gas-atomized copper
different than that of ceramics. Being good powders of different particle sizes were sup-
electrical conductors, no internal electrical field plied by American Chemet Corporation, USA.
is induced in metals. Microwave interaction The average (d50) particle sizes were 6, 12, 18,
with metals is restricted to its surface only. The 63 and 383 μm respectively. The experimental
depth of penetration in metals, also known as part consists of two subsets. The first experiment
skin depth, is defined as the distance into the investigates the effect of particle size during
material at which the incident power drops to the heating of powder particle in a multimode
1/e (36.8 %) of the surface value. In general, microwave furnace. The second experiment
the skin depth is relatively small in metals, since evaluates the role of initial density during the
in the microwave regime, the particle sizes are heating of metal powder in a multimode micro-
much smaller than the wavelength of microwave wave furnace. In the first experiment powders
radiation; the field across the particle is uniform were uniaxially compacted at 400 MPa in to
and causes volumetric heating [Takayama et al., a cylindrical pellet (12.7 mm diameter and 5
2006]. However, for relatively coarse particle mm average height) of green densities ranging
(>100μm), the heating may be conductive from between 60 to 83% of the theoretical density
outside to the interior of the powder. According (8.9 g/cm3). For the second part of the experi-
to the Faraday’s effect in a conductive material, ment, cylindrical pellets of the same dimensions
a varying magnetic field generates an electric of as-received gas-atomized copper powders
field that gives rise to eddy currents and sub- (particle size 18 μm) were compacted using the
sequent resistive losses. Hence a metal powder same compaction press at 90 to 350 MPa in to
compact will start absorbing more microwave a green densities ranging between 56 to 76%
radiation with an increase in temperature due to of the theoretical density. Microwave sintering

43-1-6 Journal of Microwave Power & Electromagnetic Energy ONLINE Vol. 43, No. 1, 2009
Figure 2. Thermal profiles of copper compacts as a function of particle size.

temperature. However, as very little variation in In the above expression, σ and ω are the DC
the emissivity was reported in the temperature electrical conductivity and angular frequency,
range used in the present study, hence, the ef- respectively. At 2.45 GHz, the skin depth of bulk
fect of variation in emissivity was ignored in the copper is about 1.3μm. But, as in case of porous
present investigation. material, electrical conductivity decreases with
an increase in porosity. Therefore, based on the
RESULT AND DISCUSSION above equation, effective skin depth should also
increase. So the difference in the heating rates
Figure 2 compares the thermal profiles of the can be attributed to the difference in skin depth
copper powder compacts of varying particle size, and also the change in surface area per unit vol-
sintered in multimode microwave furnace. It is ume with the change of particle size. The effect
interesting to note that the porous metal compact is more pronounced for the larger size powders
couple with microwaves heats rapidly. As par- because they do not retain their single-particle
ticle size increases the heating rate decreases and status when pressed into a compact and are
after certain time heating rate becomes constant more likely to form electrical contacts during
at a particular power setting. It is well known that die pressing, making them effectively much
microwaves for electrically conducting materials larger. This was experimentally proved by Ma
such as copper do not penetrate a bulk sample et al. [2007]. They measured the conductivity of
beyond the skin depth. The skin depth is given a green sample made from 22 μm powder and
by the well-known expression: found that it was 104 times as great as one made
from 3 μm powder.
 Another factor that influences the heat-
 
   ing behavior is the initial green density of the
metal powder compacts. Figure 3 shows the
Here, μ is the real part of the permeability taken experimental temperature rise with time in pure
to be 4 × 10-7 Tm/A for nonmagnetic materials. copper compacts with 56%, 65%, 71% and 76%

43-1-8 Journal of Microwave Power & Electromagnetic Energy ONLINE Vol. 43, No. 1, 2009
and D. T. Zimmermana (2007). “ Systematic Study of
Microwave Absorption, Heating and Microstructure
Evolution of Porous Copper Powder Metal Com-
pacts.” J. Appl. Phys., 101, pp.074906-08.
Mishra, P., G. Sethi, and A. Upadhyaya (2006). “Model-
ing of Microwave Heating of Particulate Metals.”
Metallurgical and Materials Transactions B, 37(b),
pp.839-845.
Nayer, P. (1997). The Metals Data Book. McGraw-Hill,
New York.
Pert, E., Y. Carmel, A. Birnboim, T. Olorunyolemi, D.
Gershon, J. Calame, I.K. Lloyd, and O.C. Wilson Jr.
(2001). “Temperature Measurement during Micro-
wave Processing: The significance of Thermocouple
Effects.” J. Am. Ceram. Soc., 84(9), pp.1981-1986.
Porada, M.W., and R. Borchert (1996-1996). “Microwave
Sintering of Metal-Ceramic FGM.” Proceedings of
the 4th International Symposium on Functionally
Graded Materials, AIST Tsukuba Research Center,
Tsukuba, Japan, October 21- 24, pp. 349-354.
Rodiger, K., K. Dreyer, K. Gerdes, and M.W. Porada
(1998). “Microwave Sintering of Hardmetals.” In-
ternational Journal of Refractory Metals & Hard
Materials, 16, pp. 409-416.
Roy, R., D.K. Agrawal, J.P. Cheng, and S. Gedevanishvili
(1999). “Full Sintering of Powdered Metals Using
Microwaves.” Nature, 399(17), pp.668-670.
Rybakov, K.I., V. E. Semenov, S. V. Egorov, A. G. Ere-
meev, I. V. Plotnikov, and Yu. V. Bykov (2006). “Mi-
crowave Heating of Conductive Powder Materials.”
J. Appl. Phys., 99, 023506-09.
Saitou, K. (2006). “Microwave Sintering of Iron, Cobalt,
Nickel, Copper and Stainless Steel Powders.” Scripta
Materialia, 54(5), pp.875-879.
Sethi, G., A. Upadhyaya, and D. Agrawal (2003). “Micro-
wave and Conventional Sintering of Premixed and
Pre-alloyed Cu-12Sn Bronze.” Science of Sintering,
35, pp.49-65.
Takayama, S., G. Link, S. Miksch, M. Sato, J. Ichikawa,
and M. Thumm (2006). “Millimetre Wave Effects
on Sintering Behavior of Metal Powder Compacts.”
Powder Metallurgy, 49(3), pp.274-280.
Vaidhyanathan, B. and K.J. Rao (1997). “Microwave
Assisted Synthesis of Technologically Important
Transition Metal Silicides.” J. Mater. Res., 12(12),
pp.3225-29.

43-1-10 Journal of Microwave Power & Electromagnetic Energy ONLINE Vol. 43, No. 1, 2009
MICROWAVE SINTERING OF W-18CU AND
W-7NI-3CU ALLOYS
Avijit Mondal1*, Anish Upadhyaya1 and Dinesh Agrawal2
Department of Materials & Metallurgical Engineering
1

Indian Institute of Technology, Kanpur 208016, India


2
The Pennsylvania State University, University Park, PA 16802, USA
*
avijitm@iitk.ac.in

Microwave processing is emerging as an innovative and highly effective material processing


method offering many advantages over conventional methods, especially for sintering applications.
It is recognized for its various advantages, such as: time and energy saving, rapid heating rates,
considerably reduced processing cycle time and temperature, fine microstructures and improved
mechanical properties which lead to better product performance. Major constraints in conventional
sintering of refractory material such as tungsten and its alloys are high sintering temperatures and
long soaking times which cause abnormal grain growth and lead to poor mechanical properties.
They get further aggravated at smaller (submicron and nano) tungsten powder sizes. This study
describes recent research findings; W-18Cu and W-7Ni-3Cu alloys have been successfully con-
solidated using microwave heating which resulted in an overall reduction of sintering time of up
to 80%. The microwave sintered samples exhibited finer microstructure and superior mechanical
properties when compared with the conventional samples.
Submission Date: 6 August 2008
Acceptance Date: 7 January 2009
Publication Date: 16 January 2009

INTRODUCTION processed via sintering at temperatures ranging


from 1200°C to 1500°C when at least one of the
Tungsten is a refractory metal with a melting constituents melts and facilitates the sintering
point of 3420°C. Because of its high melting process.
point, tungsten and its alloys are mostly consoli- Depending upon the composition, these
dated through powder metallurgy (P/M) tech- alloys also offer a favorable combination of
niques, though for some specific applications high thermal conductivity, low thermal expan-
mechanical alloying and infiltration techniques sion coefficient, high density, high strength and
are also employed. Most commonly, liquid phase ductility, good corrosion resistance and ease of
sintering (LPS) is used for consolidating these machinability. These alloys, because of their
alloys. The technique offers the advantage of rel- wide range of properties, are strategically very
atively lower sintering temperatures, enhanced important and have a wide range of applications,
densification, microstructural homogenization such as electrical contact, heat spreaders, elec-
and near theoretical density. These alloys are tronic packaging, radiation-shield, counter-bal-
anced weights, and kinetic energy penetrator in
Keywords: Tungsten heavy alloys, microwave sintering, defense industry [Upadhyaya, 2001]. Because of
microstrutures technological demands and strategic concerns,
there is considerable interest worldwide in con-
Guest Editor: Dr. Satoshi Horikoshi, Tokyo University of
Science, Chiba, Japan
solidating the tungsten based alloys for thermal

International Microwave Power Institute 43-1-11


management and defense industry applications. in microwaves [Saitou, 2006; Jain et al., 2006;
Each application has its own specific require- Upadhyaya et al., 2008; Mondal et al., 2008].
ments as far as the properties are concerned. The present study describes the consolidation of
In general for most of the applications, near tungsten based alloys using microwave energy.
theoretical density, dimensional stability, high Conventional sintering has also been conducted
hardness, toughness and very high ductility are to compare the effectiveness of the microwave
important [German, 1996]. process.
In order to avoid thermal shock, processing
of tungsten heavy alloys in a conventional fur- EXPERIMENTAL PROCEDURE
nace involves heating at a slower rate (<10°C/
min) and with an isothermal hold at intermittent For the present study the W-18Cu powders were
temperatures. This not only increases the process supplied by Osram Sylvania, Towanda, USA.
time, but also results in significant microstruc- The elemental W (d50: 4.2μm) , Ni (d50: 11.0 μm)
tural coarsening (W grain growth) during sinter- and Cu (d50: 47.0 μm) powders were supplied
ing, leading to the degradation of mechanical by Kennametal-Widia India Ltd., Bangalore;
properties. This problem is further aggravated International Specialty Products (INCO), USA;
when the initial powder size is extremely fine. and ACu Powder International, LLC, USA re-
Hence, it is envisaged that a fast heating rate spectively. The detailed powder characteristics
would mitigate this problem. One of the tech- have been reported elsewhere [Mondal et al.,
niques to achieve fast and relatively uniform 2008]. The elemental W, Ni and Cu powders
sintering is through microwaves. Microwaves were mixed in a required proportion in a turbula
interact with individual particulates within the mixer for about 60 min to prepare 90W-7Ni-3Cu
pressed compacts directly, and thereby provide (wt.%) alloy composition. Powders were pressed
rapid volumetric and uniform heating [Clark in a die of 16 mm inner diameter to make the
and Sutton, 1996]. This technique offers many green compacts of approximately 6 to 8 mm in
advantages over conventional methods for con- height. The pressure was applied uniaxially in a
solidation. The main advantages include rapid 50T hydraulic press with floating die, using zinc
sintering, fine microstructure, energy savings stearate as a die wall lubricant. All the samples
and improvements in the mechanical properties were compressed at pressures of 200 MPa.
of the sintered products. Until recently, micro- To study the densification behavior, the green
wave processing was mostly restricted to ceram- compacts were sintered using conventional and
ics, cemented carbides and ferrites [Clark and microwave furnaces. The sintering temperatures
Sutton, 1996; Gerdes et al., 1996; Agrawal et al., selected for the liquid phase sintering of W-18Cu
2001; Rodiger et al., 1998; Porada and Borchert, and 90W-7Ni-3Cu were 1300°C and 1450°C,
1996-97; Tsay et al., 2001]. Applicability of respectively. Sintering was carried out with a
microwave sintering to metals was ignored due constant heating rate 5°C/min in a flowing H2
to the fact that they reflect microwaves. Roy et atmosphere. Microwave sintering of the green
al. [1999] reported that particulate metals can compacts was carried out using a multi-mode
be heated rapidly in microwaves. This has led cavity 2.45 GHz, 6 kW commercial microwave
to the use of microwaves to consolidate a range furnace. The temperature of the sample was
of particulate metals and alloys. Researchers monitored using an optical pyrometer with the
have reported microstructural refinement due circular crosswire focused on the sample cross-
to rapid heating, significant reduction in pro- section.
cess time, elimination of brittle intermetallic The sintered density was obtained by both di-
formation and superior mechanical properties mensional measurements as well as Archimedes’

43-1-12 Journal of Microwave Power & Electromagnetic Energy ONLINE Vol. 43, No. 1, 2009
(a) (b)
Figure 1. Thermal profiles of (a) W-18Cu and (b) W-7Ni-3Cu alloys sintered using conventional
and microwave furnaces.

density measurement technique. Metallographic RESULTS AND DISCUSSION


techniques were employed on the sintered
samples. The scanning electron micrographs of Figs. 1(a) and 1(b) compare the thermal profiles
as-polished samples were obtained by a scanning for both the compositions heated in both the
electron microscope (model: FEI quanta, Neth- conventional and microwave furnace. It is in-
erlands) in the secondary electron (SE) and back teresting to note that for both, the compositions
scattered (BSE) mode. Phase determination and couple with microwaves and heat up rapidly.
phase evolution, if any were studied for all the The overall heating rates in the microwave fur-
samples with the help of X-Ray Diffractometer nace were 15° and 22°C/min for W-18Cu and
(Model: RICH. SEIFERT & Co., GmbH & Co. 90W-7Ni-3Cu, respectively. Taking into con-
KG, Germany). The experimental variables of sideration the slow heating rate (5°C/min) and
the XRD are as follows: scan rate-3°/min, tar- isothermal holds at intermittent temperatures in
get-Cu, and power-30kVx20 mA. Bulk hardness conventional furnace, there is about a 70 to 80%
measurements were performed on polished sur- reduction in the process time during sintering of
faces of sintered cylindrical compacts at a load all the compositions in the microwave furnace.
of 5 kg using Vickers hardness tester (model: Despite such a fast heating rate, no micro- or
V100-C1, supplier: Leco, Japan). The load was macro-cracking was observed in all microwave-
applied for 30 s. Micro-hardness tester (model: sintered samples. This underscores the efficacy
Leitz 8299, Germany) was used to perform the of volumetric heating associated with micro-
micro-hardness tests on the tungsten grains and waves. Figs. 2(a) and 2(b) show the effect of
matrix phases. The loads applied for micro-hard- the heating mode upon the sintered density and
ness test was 50 g for 15 s. Ten measurements densification parameter for both the composi-
were taken at random locations for each macro tions. It is noteworthy to mention that in both, a
and micro-hardness test to determine statistically heating mode more than 98% of its theoretical
accurate results. density has been achieved. Microstructural ob-
servation was made to confirm the effect of the
heating mode on grain coarsening phenomenon
and microstructural homogeneity.

International Microwave Power Institute 43-1-13


Table 1. Effect of heating mode on the hardness of W-18Cu and 90W-7Ni-3Cu compacts.
Heating mode W-18Cu 90W-7Ni-3Cu
Bulk hardness Bulk hardness Micro-hardness
W grains Matrix
Conventional 325±12 301±21 445±15 333±13
Microwave 376±15 328±12 442±9 312±18

For both the compositions, microwave sintering International Conference on Tungsten,Refractory and
results in a more refined microstructure, which Hardmaterials VII , Washington DC, USA.
Mondal, A., and A. Upadhyaya (2008). “Phase Evolution
in turn, leads to enhancement of hardness. of Tungsten Alloys during Microwave Sintering.”
unpublished research.
ACKNOWLEDGMENTS Porada, M.W., and R. Borchert (1996-1997). “ Microwave
Sintering of Metal-Ceramic FGM.” Proceedings of
This collaborative research was done as a part of the 4th International Symposium on Functionally
Graded Materials, AIST Tsukuba Research Center,
the Center for Development of Metal-Ceramic Tsukuba, Japan, October 21–24, 1996-1997, pp.
Composites through Microwave Processing, 349-354.
funded by the Indo-US Science and Technol- Rodiger, K., K. Dreyer, T. Gerdes, and M. W. Porada,
ogy Forum (IUSSTF), New Delhi. The authors (1998). “Microwave Sintering of Hardmetals.” In-
thank Dr. D. Houcke and Mr. P. Sedor of Osram ternational Journal of Refractory Metals & Hard
Materials, 16, pp.409-416.
Sylvania, Towanda, USA for supplying the Roy, R., D.K. Agrawal, J.P. Cheng, and S. Gedevanishvili
powders. (1999). “Full Sintering of Powdered Metals Using
Microwaves.” Nature, 399(17), pp.668-670.
REFERENCES Saitou, K. (2006). “Microwave Sintering of Iron, Cobalt,
Nickel, Copper and Stainless Steel Powders.” Scripta
Materialia, 54(5), pp.875-879.
Agrawal, D.K., A.J. Papworth, J. Cheng, H. Jain, and Tsay, C.Y., K.S. Liu, and I.N Lin, (2001). “Co-firing Pro-
D.B. Williams (2001). “ Microstructural Examina- cess Using Conventional and Microwave Sintering
tion by TEM of WC/Co Composites Prepared by Technologies for MnZn- and NiZn-Ferrites.” Journal
Conventional and Microwave Processes.” Proc. 15th of European Ceramic Society, 21, pp.1937-40.
International Plansee seminar, v.2, G. Kneringer, P. Upadhyaya, A. (2001). “Processing Strategy for Con-
Rodhammer and P. wilhartitz (eds.), Plansee AG, solidating Tungsten Heavy Alloys for Ordnance
Reutte, Austroa, pp.677-684. Applications.” Materials Chemistry and Physics,
Clark, D.E and W. H. Sutton (1996). “Microwave Pro- 67, pp.101-110.
cessing of Materials.” Ann. Rev. Mater. Sci., 26, pp. Upadhyaya, A., S. K. Tiwari, and P. Mishra (2007).
299-331. “Microwave Sintering of W-Ni-Fe Alloys.” Scripta
Gerdes, T., M. W. Porada, K. Rodiger, and K. Dreyer Materialia, 56, pp.5-8.
(1996). “Microwave Sintering of Tungsten Carbide
– Cobalt Hardmetals.” Mater. Res. Soc. Symp. Proc.,
430, pp.175-180.
German, R.M. (1996). Sintering Theory and Practice. J.
Wiley, New York.
Jain, M., G. Skandan, K. Martin, K. Cho, B. Klotz, R.
Dowding, D. Kapoor, D. Agarwal, and J. Chang
(2006). “Microwave Sintering: A New Approach to
Fine-Grain Tungsten-I.” International Journal of
Powder Metallurgy, 42(2), pp.45-50.
Mondal, A., D. Agarwal, and A. Upadhyaya (2008).
“Microwave Sintering of Tungsten Based Alloys.”

43-1-16 Journal of Microwave Power & Electromagnetic Energy ONLINE Vol. 43, No. 1, 2009
Science of Sintering, 42 (2010) 99-124
________________________________________________________________________

doi: 10.2298/SOS1001099S

UDK 675.92.027:54-732
Numerical Modeling of Microwave Heating

A. K. Shukla, A. Mondal, A. Upadhyaya*)


Department of Materials & Metallurgical Engineering
Indian Institute of Technology, Kanpur 208016, India

Abstract:
The present study compares the temperature distribution within cylindrical samples
heated in microwave furnace with those achieved in radiatively-heated (conventional)
furnace. Using a two-dimensional finite difference approach the thermal profiles were
simulated for cylinders of varying radii (0.65, 6.5, and 65 cm) and physical properties. The
influence of susceptor-assisted microwave heating was also modeled for the same. The
simulation results reveal differences in the heating behavior of samples in microwaves. The
efficacy of microwave heating depends on the sample size and its thermal conductivity.
Keywords: Microwave heating; Heat Transfer, Finite Difference Modelling (FDM)

1. Introduction

Microwaves are now well-recognized mode of heating material. In the beginning (late
40’s) 2.45 GHz microwaves were first used for heating food products up to 100-150°C [1].
Subsequently, they were used in rubber industry for aiding vulcanization [1]. Later on,
microwaves were utilized for heating ceramic bodies [2]. It was reported that microwaves
interact both with bulk as well as particulate ceramics and can cause rapid increase in their
temperature to up to 1700°C [3]. The coupling of particulate ceramics elicited keen interest
among researchers as it provided means to rapidly sinter the ceramic powder compacts [4]
without significant microstructural coarsening. Unlike conventional heating, microwaves
directly couple with the material, which leads to rapid and uniform heating. In the context of
sintering this is particularly important. Over the years, a variety of materials such as carbides,
nitrides, complex oxides, silicides, and apatites, etc. have been synthesized using microwaves.
Despite the widespread application of microwaves for thermal processing of materials, their
use till recently restricted for metals. This was on account of the fact that bulk metals tend to
reflect most of the microwaves incident on them, and only heat over a localized surface. Such
an interaction frequently results in arcing that eventually causes the failure of the magnetron.
However, in 1999, for the first time Roy et al. [5] demonstrated that most of the particulate
metals very effectively couple with microwaves and can be heated to high temperatures at
significantly faster rates. This has opened the possibility of consolidating particulate metal
compacts to full density through microwave sintering [6]. Several researchers [7-16] in recent
years have demonstrated the efficacy of microwave sintering in a range of materials that
include Cu, Sn, bronzes, stainless steel and tungsten heavy alloys.
Despite the widespread use of microwaves for heating materials, there is still no
unanimity regarding its interaction mechanism with the materials, particularly, metallic ones
_____________________________
*)
Corresponding author: anishu@iitk.ac.in
100 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

[17-21]. The main loss mechanisms during microwave-material interaction are electric,
conduction (eddy current), hysteresis and resonance (domain wall and electron spin (FMR)).
It is often difficult to ascertain which loss mechanism, or combination of mechanisms is
occurring for a particular sample in given conditions. The different mechanisms do however
have different dependencies on certain properties such as sample type and microstructure,
frequency and temperature.
Recently, Mishra et al. [22] modeled the heating of particulate metal compacts by
solving the Maxwell equation and using a 2D-Finite Difference Time Domain (FDTD) model.
While their model showed good correlation between the experimental and simulated
temperature profiles of Cu, Sn and W, it had several limitations. Their model ignored the
effect of compact size and its thermal conductivity and assumed a uniform temperature
distribution within the compact. Furthermore, it failed to compare the thermal profile of
microwave heated compact with those achieved using a conventional (radiatively heated)
furnace. In case of microwave heating, since the sample per se acts as the source of heat,
hence, it is hypothesized that the compact should have an inverse thermal profile with the
central (inside) portion having higher temperature than the peripheral regions. Furthermore, at
times, a low temperature the coupling between microwaves and the compact is poor. To
address these issues, in practice, microwave heating is usually done by surrounding the
compact with susceptors such as SiC that gets preferentially heated at lower temperatures.
The present study, therefore systematically investigates the evolution of temperature
profile with time within a cylindrical material as a function of its size (0.65 cm, 6.5 cm, 65
cm) and conductivity under both conventional as well as microwave heating mode. The latter
has been examined both in the presence and absence of susceptor. The objective of current
study is to device a well defined strategy for the selection of most effective sintering process
(conventional versus microwave) for the given conditions. To achieve this, 2-D finite
difference mathematical model has been formulated to examine the role of various process
variables.

2. Mathematical Formulation

For the present simulation, the material was considered to interact with 2.45 GHz
multimode microwaves. Fig. 1 shows the schematic of a sample configuration in a multimode
microwave furnace. For facilitating the modeling an axio-symmetric (cylindrical) sample
geometry was assumed with the radius (R) to height (H) ratio of unity. Three different radii
(0.65, 6.5 and 65 cm) cylinders were considered in the present study for simulation. To
simulate the effect of thermal conductivity on the heating behavior for each size the
simulation was performed for a low and high thermal conductivity condition, respectively.
The present configuration was constructed so as to enable the placement of SiC
susceptors for assisting microwave heating. Microwaves penetrate and propagate through a
dielectric material like SiC. This generates an internal electric field (E) within a specific
volume, which in turn, induces polarization and movement of charges. The resistance to these
induced motions due to internal, elastic and frictional forces attenuates the electric field.
These losses result in volumetric heating. The resulting electromagnetic power absorbed per
unit volume (PEM) [Wm-3] by a material is given by [23]:

PEM = 2ω {εoεr′′ (T) |Erms|2 + µoµr′′ (T) |Hrms|2} (1)


where,
ω is frequency of the microwave radiation (2.45 GHz)
ε0 is absolute permittivity of free space (8.85×10-12 C2N-1m-2)
εr′′ is relative imaginary component of the dielectric constant
µ0 is absolute permeability of free space (4π×10-7 NA-2)
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 101
___________________________________________________________________________

µr′′ is relative imaginary component of permeability constant


Erms and Hrms are root mean square values of electric and magnetic field amplitudes,
respectively.

Fig. 1. Schematic of the microwave setup used in the present study.

This heat generation rate per unit volume in the material during microwave sintering
can be designated as the source term, S. For the sake of simplification, for constant power and
frequency input, the magnetic permeability, electrical conductivity and permittivity are
assumed to be linear function of temperature. Elsewhere too, this approximation has been
considered [24]. The entire source term ‘S’ therefore can be expressed as:
S = ASOURCE + BSOURCE ×T (2)
where, ASOURCE and BSOURCE are the constants and are specific to the materials and microwave
operating parameters (frequency and power).
The differences between conventional and microwave heating can be described
mathematically using unsteady state heat transfer equation with a volumetrically distributed
heat source and is expressed as:
∂T 1 ∂ ⎛ ∂T ⎞ ∂ ⎛ ∂T ⎞
ρc p = ⎜ kr ⎟ + ⎜k ⎟+S (3)
∂t r ∂r ⎝ ∂r ⎠ ∂z ⎝ ∂z ⎠
where, S is local density of heat sources. For conventional heating ‘S’ is non-existent,
whereas for microwave processing it can be determined using the Equations 1 and 2. Having
calculated the source term, the heat transfer for the cylindrical geometry was done using a 2D
finite difference explicit model. Fig. 2 shows the flowchart of the entire modeling scheme.
For formulating the model, the following assumptions have been made
(i) A cylindrical coordinate system has been assumed.
(ii) Physical properties, such as density, specific heat, thermal conductivity, resistivity
are assumed to be uniform throughout the calculation domain.
(iii) Heat source is uniformly distributed.
102 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 2. Flow diagram of the 2-D finite difference model used for cylindrical geometry.

Rearranging the governing heat transfer Equation 3:


∂T
ρc p
∂t = ∂ T + 1 ∂T + ∂ T + S
2 2
(4)
k ∂r 2 r ∂r ∂z 2 k

For the finite difference modeling, the cylindrical sample geometry is subdivided into
radial and axial grids. The total number of grid points in the radial and axial directions has
been designated as M and N, respectively (Fig. 3). The grid spacing in both the directions, Δr
and Δz, can be expressed as following:
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 103
___________________________________________________________________________

R
Δr = (5)
( M − 1)

H
Δz = (6)
( N − 1)

Fig. 3. (a) Discretization of the calculation domain of the cylindrical compact having the
radius to height ratio of unity. (b) Highlighted area in represents the section for which axial
and radial temperature profiles and isothermal contours were calculated.

The temperature within the sample at any grid point (m,n) at a specific time step (l) can be
designated as T (m,n,l), where
m = radial grid points 1, 2, 3, 4…..M (7.1)
n = axial grid points 1, 2, 3, 4…..N (7.2)
l= time step number 1, 2, 3… max t (7.3)
After discretizing, Equation (4) can be expressed as:
1 T (m, n, l + 1) − T (m, n, l ) T (m + 1, n, l ) − 2T (m, n, l ) + T (m − 1, n, l )
⋅ =
α Δt (Δr )2 (8)
1 T (m + 1, n, l ) − T (m, n, l ) (T (m, n + 1, l ) + T (m, n − 1, l ) − 2T (m, n, l )) S
+ + +
(m − 1)(Δr ) (Δr ) (Δz ) 2 k

After rearranging, the expression becomes

α .Δt T (m + 1, n, l ) − 2T (m, n, l ) + T (m − 1, n, l ) α .Δt ⎡ T (m + 1, n, l ) − T (m, n, l ) ⎤


T (m, n, l + 1) = +
(Δr ) 2 (Δr ) 2 ⎢⎣ m −1 ⎥
⎦ (9)
α .Δt (T (m, n + 1, l ) + T (m, n − 1, l ) − 2T (m, n, l )) S k
+ + . + T (m, n, l )
(Δz ) 2 k ρc p
The above equation is valid only for the inner nodes {m: 1, 2, 3 …..M-1 and n:
2…..N-1}. For writing the finite difference equation for the nodes lying on the outer surface
and edges the equations have to be reformulated. Fig.s 4a and 4b show the section view of
side and top/bottom surface, respectively of the cylindrical geometry. For the various
locations of the cylinder the nodal points can be expressed as following:
Side Surface m = M; n = 2, 3……….N-1 (10.1)
Top & Bottom Edges m = M; n = 1,N (10.2)
Top & Bottom Surface M = 1,2…..M; n = 1,N (10.3)
104 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 4. Section of the cylindrical compact showing (a) the side surface and (b) top/bottom
surface node.

2.1. Heat Transfer at Side Surface Boundary

In Fig. 4a, the heat balance in control volume centered at location corresponding to
node ‘O’ can be expressed as the summation of the rate of heat flow from various adjacent
nodes as following:

Rate of heat accumulation in control volume at the side surface



Rate of heat flow through conduction from {node 1→ node O + 2→O + 4→O + 5→O +
6→O}
+
Rate of heat flow through convection from 3→O
+
Rate of heat flow through radiation from 3→O
+
Rate of heat flow through susceptor 3→O
+
Rate of heat generation in control volume (11)

The above expression can be expressed mathematically as following:


A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 105
___________________________________________________________________________

{k .(Δθ .Δz )( R − Δr ). T (m − 1, n, l ) − T (m, n, l ) + h.(Δθ .Δz ).R.(Tamb − T (m, n, l ))


2 Δr
Δr T (m, n − 1, l ) − T (m, n, l )
+ k SiC .(Δθ .Δz ).R.(TSiC − T (m, n, l )) + k .(Δθ . ).R.
2 Δz
(12)
Δr T (m, n + 1, l ) − T (m, n, l )
+ k .(Δθ . ).R. + σ .ε .F .(Δθ .Δz ).R.(Tend
4
− T 4 (m, n, l ))
2 Δz
Δr ⎫ Δr
+ S .( R.Δθ ).Δz. ⎬Δt = ρc p . .( R.Δθ ).Δz.[T (m, n, l + 1) − T (m, n,.l )]
2⎭ 2
Rearranging Equation 12,
2 k .Δt ⎡ Δr T (m − 1, n, l ) − T (m, n, l ) ⎤ h (Tamb − T (m, n, l ))
T ( m, n, l + 1) = T (m, n, l ) + . ⎢( R − ) ⎥ + .R.
r ρc p ⎣ 2 (Δr )2 ⎦ k (Δr )
k SiC (T 4 − T 4 ( m, n, l )) σ .ε .F (T 4 − T 4 (m, n, l )) (T (m, n − 1, l ) − T (m, n, l ))
+ .R. SiC + .R. end + R.
k (Δr ) k (Δr ) 2(Δz ) 2
(T (m, n + 1, l ) − T (m, n, l )) S .Δt
+ R. +
2(Δz ) 2 ρc p
(13)

2.2 Heat Transfer at Top/Bottom Edge Boundary

As done with the surface nodes, the heat balance in control volume centered at the top
and the bottom edges (Fig. 4b) location corresponding to node ‘O’ can be expressed as the
summation of the rate of heat flow from various adjacent nodes to it as following:

Rate of heat accumulation in control volume at top/bottom edge



Rate of heat flow through conduction from {1→O + 2→O + 4→O + 6→O}
+
Rate of heat flow through convection from {3→O + 5→O}
+
Rate of heat flow through radiation from {3→O + 5→O}
+
Rate of heat flow through susceptor {3→O + 5→O}
+
Rate of heat generation in control volume (14)

The above expression is mathematically expressed as:

4 k .Δt ⎡1 Δr T ( m − 1, n, l ) − T ( m, n, l ) ⎤ h (Tamb − T (m, n, l )) 1 ⎡ 1 1⎤


T (m, n, l + 1) = T ( m, n, l ) + . ⎢ (R − ) ⎥ + .R. . .⎢ + ⎥
r ρc p ⎣ 2 2 (Δ r )2
⎦ k 2 ⎣ Δ r Δ z⎦
k SiC (T − T ( m, n, l )) 1 ⎡ 1 1 ⎤ σ .ε .F (T 4 − T 4 (m, n, l )) 1 ⎡ 1 1⎤
+ .R. SiC . .⎢ + ⎥ + .R. end . .⎢ + ⎥
k ( Δr ) 2 ⎣ Δr Δz ⎦ k ( Δr ) 2 ⎣ Δr Δz ⎦
(T ( m, n + 1, l ) − T (m, n, l )) (T (m, n − 1, l ) − T ( m, n, l )) S .Δt
+ R. + R. +
2( Δz ) 2 2 ( Δz ) 2 ρc p
(15)
106 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

2.3. Heat Transfer at Top/Bottom Surface Boundary

Rate of heat accumulation in control volume at top/bottom surface



Rate of heat flow through conduction from {1→O + 2→O + 3→O + 4→O + 6→O}
+
Rate of heat flow through convection from {5→O}
+
Rate of heat flow through radiation from {5→O}
+
Rate of heat flow through susceptor {5→O}
+
Rate of heat generation in control volume (16)

The above expression can be mathematically formulated as:


{k .(Δθ . Δz )(r − Δr ) T (m − 1, n, l ) − T (m, n, l ) + h.(Δθ .Δr ).r.(Tamb − T (m, n, l ))
2 2 Δr
T (m, n − 1, l ) − T (m, n, l ) T (m, n + 1, l ) − T (m, n, l )
+ k .(Δθ .Δr ).r. + k .(Δθ .Δr ).r.
Δz Δz (17)
+ k SiC .(Δθ .Δr ).r.(TSiC − T (m, n, l )) + σ .ε .F .(Δθ .Δr ).r.(Tend − T (m, n, l ))
4 4

Δz ⎫ Δz
+ S .(r.Δθ ).Δr. ⎬Δt = ρc p . .(r.Δθ ).Δr.[T (m, n, l + 1) − T (m, n, l )]
2⎭ 2
Rearranging the above,
2 k .Δt ⎡ Δr T (m − 1, n, l ) − T (m, n, l ) ⎤ h (Tamb − T (m, n, l ))
T (m, n, l + 1) = T (m, n, l ) + . ⎢( r − ) ⎥ + .r.
r ρc p ⎣ 2 (Δr )2 ⎦ k Δz
k (T − T (m, n, l )) σ .ε .F (Tend − T (m, n, l ))
4 4 (18)
+ SiC .r. SiC + .r.
k (Δz ) k (Δz )
(T (m, n − 1, l ) − T (m, n, l )) (T (m, n + 1, l ) − T (m, n, l )) S .Δt
+ r. + r. +
2(Δz ) 2 2(Δz ) 2 ρc p

2.4. Error Analysis and Establishing Stability Criteria

The finite difference model involves numerically solving the equation using Taylor
series expansion and hence consists of some inherent errors [25]. One such error is the
truncation error which can be attributed to the computation of derivatives that are used in
place of the finite difference equations. Another prevalent error in such formulation is the
numerical round-off error that is a consequence of finite significant digit restriction [26]. For
the explicit forward Euler method used in the present study, it is therefore pertinent to
establish the stability criterion. The time step, Δt, for the temporal temperature simulation
within the sample is expressed as [27]:
αΔt 1
Fo = ≤ (19.1)
1 2
(Δr 2 + Δz 2 )
2

After rearranging, the above equation Δt can be expressed as


A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 107
___________________________________________________________________________

Δr 2 Δz 2
Δt ≤ + (19.2)
8α 8α
The stability criterion calculated in the above equation is valid only for the inner nodes.
Similarly, calculations can be made for the boundary nodes as well and are expressed in terms
of Biot number (Bi) as following:
1
heff . (Δz + Δr )
Bi = 2 (20.1)
k
where,
heff = h + k SiC + σε (TSiC
2
+ Tamb
2
).(TSiC + Tamb ) (20.2)

1
Fo ≤ (20.3)
2.(2 + Bi )
For the present simulation, assuming a 50×50 (M,N) mesh, irrespective of the sample size and
its conductivity Δt ≤ 0.0015s was found to satisfy the stability criterion.

2.5. Calculation Procedure

Tab. I provides a summary of the physical properties and other parameters used as
inputs for the simulation. To obtain the thermal profiles within the compact at various time
intervals, the mathematical formulation of the heat transfer model was done using a
FORTRAN based computer program run under UNIX environment. The simulation was run
for 180 seconds time duration for both conventional as well as microwave heating. The result
output from the compiled program was in the form of a three-dimensional matrix providing
the temperature at each node within the compact at various time steps. The data from the
output files were plotted using MATLAB (ver. 2008a).

Tab. I. Summary of the physical properties and other parameters used as inputs for
thesimulation.

Density (Kgm-3) 5000


Specific heat (Jkg-1 K-1) 400
Heat Transfer Coefficient (Wm-2 K-1) 50
100
Susceptor Constant (Wm-2 K-1) 40 (high) and 1(low)
Thermal Conductivity (Wm-1K-1) 3.0 × 106
Heat source coefficient, ASource (Wm-3)
Heat source coefficient ,BSource (Wm-3 K-1) 1.5× 104
Number of grids in radial direction (M) 50
Number of grids in axial direction (N) 50

3. Results and Discussion

Fig. 5 compares the effect of heating mode on the axial and radial thermal profile at
various time steps for a high thermal conductivity cylindrical compact having radius of 0.65
cm. Fig.s 6a to 6c show the corresponding isothermal contours within cylindrical section half
108 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

(Fig. 3b) for conventional heating, only microwave heating, and susceptor-assisted
microwave heating, respectively. Tab. II summarizes the maximum temperature predicted
from our simulation for various combinations of heating, compact size and its conductivity.

Tab. II. Effect of compact size and its conductivity on the maximum predicted temperature in
conventional and microwave heating. The latter was simulated both with and without
susceptor.

Total time for


Compact Size Thermal Conductivity
Heating Mode Tmax, K Tmax calculation
Conventional 1200 1.8
High k MW without susceptor 434 180
MW with susceptor 906 180
0.65 cm
Conventional 1200 18
Low k MW without susceptor 479 180
MW with susceptor 978 180
Conventional 1200 180
High k MW without susceptor 1198 180
MW with susceptor 1523 180
6.5 cm Conventional 1200 180
Low k MW without susceptor 1597 180
MW with susceptor 1876 180
Conventional 1200 180
High k MW without susceptor 1695 180
MW with susceptor 1787 180
65 cm Conventional 1200 180
Low k MW without susceptor 1700 180
MW with susceptor 1761 180

It is worth noticing that for high conductivity small-sized compacts, at the very initial
stages (t=0.15s, 0.30s), the temperature at the center of the sample is lower than that on the
outside surface (top, bottom and side surface) (Fig. 5a). For conventional heating, the
simulations were performed for ambient surface temperature of 1200K. It is worth noticing
that for such small samples having high conductivity the temperature soon attains a steady
state condition in just 1.8s (Fig. 6a). In comparison with conventional heating, during
microwave heating the heat generation occurs within the sample per se. However, due to high
conductivity of the sample, the temperature is uniform both across the radial and axial
directions. However, even after 180s the maximum temperature attained within the compact is
below 1200K. The only difference observed is that in the presence of the susceptor the
maximum temperature attained is more than twice as much as that observed for similar
timeframe when the compacts are heated only in microwaves.
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 109
___________________________________________________________________________

Fig. 5. Simulated axial and radial temperature distribution in a high thermal conductivity
cylindrical compact having 0.65 cm radius under (a) conventional and MW heating without
(b) and with (c) susceptor assistance.

The effect of heating mode on the simulated thermal profile and isothermal contours
of a relatively low conductivity cylindrical compact to 0.65 cm radius is shown in Fig.s 7 and
8, respectively. As compared to the high conductivity counterparts, for similar dimensions,
the low conductivity samples when heated conventionally (Figs. 7a and 8a) show more
pronounced thermal gradient for the first 10s. Consequently, it takes more time for the
temperature to attain steady state condition. In contrast, for microwave heating (Fig. 7b) the
samples temperature is much more uniform (Fig. 8b). The trend remains the same even in the
presence of the susceptor. However, the maximum temperature attained is significantly higher
in susceptor-assisted microwave heating (Figs. 7c and 8c).
110 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 6. Simulated temperature distribution showing isothermal contours as a function of time


in a high conductivity cylindrical compact having 0.65 cm radius under conventional heating
(a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 111
___________________________________________________________________________

Fig. 7. Simulated axial and radial temperature distribution in a low thermal conductivity
cylindrical compact having 0.65 cm radius under (a) conventional and MW heating without
(b) and with (c) susceptor assistance.

Figs 9 and 10 show the effect of heating mode on the temporal thermal profile and
contours for high conductivity compacts of 6.5 cm radius. In spite of the high conductivity,
because of its relatively large size, conventional heating reveals steep temperature gradient in
the beginning (Figs. 9a,10a). Consequently, to homogenize the temperature distribution one
has to heat it for a longer duration. In actual systems, both these situations are detrimental to
the properties [5]. In case of microwave heating, the temperature is evenly distributed across
the section (Figs. 9b,10b). However, it does take about 180s for the temperature to reach
1200K. The same samples when subjected to microwaves in presence of SiC reveal very
interesting trend (Fig. 9c). As evident from Table II and Fig. 10c the maximum temperature
attained in the presence of the susceptor is much higher and even exceed that attained through
conventional heating. This underscores the efficacy of susceptor-aided microwave heating.
112 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Interestingly, initially, the susceptor couples more strongly than the sample. As a result the
outside surface of the sample has a higher temperature. With time however the contribution of
the same which too acts as a heat source becomes predominant. As a result at 180s, the
temperature profile becomes inverse with the sample temperature at the center being higher
than that attained outside.

Fig. 8. Simulated temperature distribution showing isothermal contours as a function of time


in a low conductivity cylindrical compact having 0.65 cm radius under conventional heating
(a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 113
___________________________________________________________________________

Fig. 9. Simulated axial and radial temperature distribution in a high thermal conductivity
cylindrical compact having 6.5 cm radius under (a) conventional and MW heating without (b)
and with (c) susceptor assistance.
114 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 10. Simulated temperature distribution showing isothermal contours as a function of time
in a high conductivity cylindrical compact having 6.5 cm radius under conventional heating
(a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 115
___________________________________________________________________________

Fig. 11. Simulated axial and radial temperature distribution in a low thermal conductivity
cylindrical compact having 6.5 cm radius under (a) conventional and MW heating without (b)
and with (c) susceptor assistance.

Fig.s 11a and 12a show the thermal profile and contours in a cylindrical compact of
6.5 cm having low conductivity heated in a conventional furnace. The thermal profile is
remarkably different than those obtained for the high conductivity samples (Fig. 9a). As
expected for low conductivity samples both the axial as well as radial thermal gradients are
very steep. While the maximum temperature is 1200K, for the timeframe used in the
simulation, the samples fail to attain a steady state condition. In contrast, for microwave
heating condition (Figs. 11b and 12b), the temperature attained is relatively uniform
throughout the samples. For longer time, the compacts do tend to get an inverse thermal
profile. Interestingly, in the presence of the susceptor the samples exhibit a mixed heating
mode. For smaller time duration the temperature at the surface is higher than that in the inside
region. For longer durations, the thermal profile becomes inverse in a more pronounced
manner.
116 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 12. Simulated temperature distribution showing isothermal contours as a function of time
in a low conductivity cylindrical compact having 6.5 cm radius under conventional heating
(a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 117
___________________________________________________________________________

Experimental results showing such a behavior has been reported by Binner and
coworkers [28]. From a design consideration, while it is recommended for medium-sized
components – particularly, those having lower conductivity – to have susceptors, the time
should be optimized so as to prevent the temperature runaways and inverse gradient.

Fig. 13. Simulated axial and radial temperature distribution in a high thermal conductivity
cylindrical compact having 65 cm radius under (a) conventional and MW heating without (b)
and with (c) susceptor assistance.

Figs 13 and 14 show the simulated temporal variation of temperatures within a high
conductivity cylindrical samples having 65 cm radius along its radius and height. Note that
despite the high conductivity, owing to the large sample size there is a thermal gradient during
conventional heating. In fact, for all the heating modes, the behavior of this sample is akin to
that shown by the medium sized sample (6.5 cm radius) having lower conductivity (Figs. 11
and 12). As shown in Fig. 15a, for such large sized cylinders, the thermal gradients are much
steeper when the samples have poor conductivity. In fact, for the same level of power setting
118 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

and simulation time, the results prove that conventional heating is virtually ineffective with
only the surface region being heated (Fig. 16a).

Fig. 14. Simulated temperature distribution showing isothermal contours as a function of time
in a high conductivity cylindrical compact having 65 cm radius under conventional heating
(a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 119
___________________________________________________________________________

This implies that for attaining the steady state condition the process time has to be
significantly increased. In case of microwave heating however, the effective thermal mass of
the sample which acts as the heat source is high. Hence, except the edges, the temperature
remains homogenous throughout the sample (Figs. 15b and 16b). In fact owing to the large
sample size, the presence of susceptor does not significantly influence the temperature
distribution within the compact (Figs. 15c and 16c). This is also reflected in the maximum
predicted temperature (Tab. II) which shows only 3.5% increase when microwave heating is
done in conjunction with susceptor.

Fig. 15. Simulated axial and radial temperature distribution in a low thermal conductivity
cylindrical compact having 65 cm radius under (a) conventional and MW heating without (b)
and with (c) susceptor assistance.
120 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

Fig. 16. Simulated temperature distribution showing isothermal contours as a function of


time in a low conductivity cylindrical compact having 65 cm radius under conventional
heating (a), MW heating without susceptor (b) and susceptor-assisted MW heating (c).
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 121
___________________________________________________________________________

Fig. 17. Effect of heating mode on the temporal thermal profile at the center of cylindrical
compact (Fig. 3b) of radius (a) 0.65 cm, (b) 6.5 cm and (c) 65 cm having high (left) and low
(right) thermal conductivity.

Fig.s 17a to 17c compare the effect of heating mode and conductivity on the
temperature variation at various time steps at the center of cylindrical compacts (Fig. 3b)
having radius of 0.65 cm, 6.5 cm and 65 cm, respectively. It is evident that irrespective of the
conductivity, for small-sized compacts (0.65 cm) conventional heating is most effective. For
medium sized compacts (6.5 cm) having higher thermal conductivity both conventional and
microwave heating results in similar maximum temperature. In the presence of the susceptor
122 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

though the temperature attained is higher. For low-conductivity medium-sized cylindrical


samples (Fig. 17b) though conventional heating is hardly effective. In comparison, microwave
(both with and without susceptor) result in significant enhancement of the sample
temperature. The most significant difference is observed for large-sized compacts (Fig. 17c)
wherein microwave are the most effective means to heat the samples in a short time span. The
above results amply demonstrate the response of varying compact size and its conductivity to
the various heating modes. It is envisaged that this will enable the design of an optimum
heating methodology customized to the physical dimensions and thermal attributes of the
specific system.

4. Conclusions

The present study investigates the effect of the sample dimension and its physical
properties on the temporal temperature evolution in microwave heating. Using a 2-D finite
difference approach the temperature distribution in cylindrical samples of varying radii (0.65,
6.5, 65 cm) and thermal conductivites was simulated for microwave as well as conventional
heating. The influence of susceptor on the sample heating behaviour in microwaves was also
modeled. In microwave heaing, it is possible to have an inverse thermal profile wherein the
temperature at the core is higher than those achieved at the surface. The tendency to exhibit
inverse thermal gradient increases with heating time and is higher for large sizes and lower
conductivity samples. Use of susceptor aids is homogenization of temperature within the
samples. For smaller-sized samples (0.65 cm) – irrespective of their thermal conductivity –
microwave heating is ineffective. In contrast, samples are uniformly heated to higher
temperatures in conventional heating in a relatively short duration. To effectively heat such
samples in microwaves susceptor is essential. For high conductivity medium-sized samples
(6.5 cm) the heating behaviour is independent of the heating mode. The beneficial effect of
microwave heating is realized for lower conductivity samples which further manifest itself in
the presence of susceptor. For much larger sample size (65 cm) – irrespective of its thermal
conductivity – microwave heating is a more practical option for attaining a uniform
temperature within a short duration. For such large sample dimensions, susceptor does not
influence the heating efficacy. Unlike microwave heating, conventional heating is restricted
only to the sample surface. This implies a more prolonged heating period which may not
always be the best option from the processing consideration. It is envisaged that the findings
from the present study can be used as a guiding tool to customize optimum heating
methodology that takes into consideration the thermal attributes and the physical dimensions
of the sample.

5. Acknowledgments

The authors thank Professor Deepak Gupta and Professor Bramha Deo of Department
of Materials and Metallurgical Engineering at IIT Kanpur for their valuable comments and
insights. The authors also gratefully acknowledge the assistance from Mr. Bijoy Sarma of
Defense Metallurgical Research Laboratory (DMRL) Hyderabad, India. This study was
supported by grants from the Indo-US Science and Technology Forum (IUSSTF) and Defense
Research and Development Organization (DRDO), India.
A. K. Shukla et al./Science of Sintering, 42 (2010) 99-124 123
___________________________________________________________________________

6. References

1. A.C. Metaxas, Foundation of Electroheat: A Unified Approach, J. Wiley & Sons


Inc., New York, NY, 1996.
2. W.R. Tinga, W.A.G. Voss, Microwave Power Engineering, E.C. Okress (ed.),
Academic Press, New York, NY, 1968, p. 189-199.
3. D.K. Agrawal, Current Opinion in Solid State Mater. Sci. 3 (1998) 480.
4. R. Roy, D.K. Agrawal, J.P. Cheng, S. Gedevanishvili, Nature 399 (1999) 668.
5. D. Agrawal, J. Mater, Education 19 (1999) 49.
6. G. Sethi, A. Upadhyaya, D. Agrawal, Sci. Sintering 35 (2003) 49.
7. R.M. Anklekar, D.K. Agrawal, R. Roy, Powder Metall. 44 (2000) 355.
8. K. Saitou, Scripta Mater. 54 (2006) 875.
9. M. Jain, G. Skandan, K. Martin, K. Cho, B. Klotz, R. Dowding, D. Kapoor, D.
Agrawal, J. Chang, Int. J. Powder Metall. 42 (2006) 45.
10. A. Upadhyaya, S.K. Tiwari, P. Mishra, Scripta Mater. 56 (2007) 5.
11. A. Mondal, D. Agrawal, A. Upadhyaya in: “Proc. of 2008 Int. Conf. on Tungsten,
Refractory & Hardmetals VII,” Metal Powder Industries Federation, Princeton, NJ,
2008, p. 3.134-3.140.
12. A. Mondal, D. Agrawal, A. Upadhyaya, Ibid., p. 3.122-3.132.
13. A. Mondal, D. Agrawal, A. Upadhyaya, Mater. Sci. Tech. 2502.
14. A. Mondal, A. Upadhyaya, D. Agrawal, J. Microwave Power Electromag. Energy
43 (2009) 11.
15. G. Prabhu, A. Chakraborty, B. Sarma, Int. J. Ref. Metals Hard Materials 27 (2009)
545.
16. A. Mondal, D. Agrawal, A. Upadhyaya, P. Chhillar, J. Cheng, R. Roy, J.
Microwave Power Electromag. Energy, 44 (2010).
17. J. Luo, C. Hunyar, I. Feher, G. Link, M. Thumm, P. Pozzo, Appl. Phys. Lett. 84
(2004) 5076.
18. S. Takayama, G. Link, S. Miksch, M. Sato, J. Ichikawa, M. Thumm, Powder
Metall. 49 (2006) 274.
19. K.I. Rybakov, V.E. Semenov, S.V. Egorov, A.G. Eremeev, I.V. Plotnikov, Y.V.
Bykov, J. Appl. Phys. 99 (2006) 023506.
20. J. Ma, J.F. Diehl, E.J. Johnson, K.R. Martin, N.M. Miskovsky, C.T. Smith, G.J.
Weisel, B.L. Weiss, D.T. Zimmerman, J. Appl. Phys. 101 (2007) 074906.
21. A. Mondal, D. Agrawal, A. Upadhyaya, J. Microwave Power Electromag. Energy
43 (2009) 5.
22. P. Mishra, G. Sethi, A. Upadhyaya, Metall. Mater. Trans. B 37B (2006) 839.
23. M. Pozar, Microwave Engineering, 2nd ed., J. Wiley & Sons Inc., Toronto, Canada,
2001.
24. J.H. Hill, T.R. Merchant, Appl. Mathematical Modelling 20 (1996) 3.
25. D.R. Poirier, G.H. Geiger, Transport Phenomena in Materials Processing, 2nd ed.,
TMS, Warrendale, PA, 1996.
26. R.J. Schilling, S.L. Harries, Applied Numerical Methods for Engineers, Thompson
Asia Pvt. Ltd. , Singapore, 2002.
27. S.C. Chapra, R.P. Canale, Numerical Methods for Engineers, 4th ed., Tata McGraw-
Hill Pub., New Delhi, India.
28. J.G.P. Binner, B. Vaidyanathan, Key Eng. Mater. 264 (2004) 725.

Садржај: У овом раду поређен је температурни распоред у цилиндричним узорцима


загреваним у микроталасној пећи са оним загреваним у радијално загрејаној
(конвенционалној). Коришћењем дводимензионог коначног приступа температурни
124 A.K. Shukla et al. /Science of Sintering, 42 (2010) 99-124
___________________________________________________________________________

профили су симулирани за цилиндре са различитим презницима (0.65, 6.5 и 65 см) и за


разлизита физичка својства. Резултати симулације указују на разлике и понашању
узорака у микроталасима. Ефикасност микроталасног загревања зависи од величине
честица и термалне проводности.
Кључне речи: Микроталасно загревање, пренос топлоте, моделирање коначном
разликом.
Hindawi Publishing Corporation
Indian Journal of Materials Science
Volume 2013, Article ID 603791, 7 pages
http://dx.doi.org/10.1155/2013/603791

Research Article
Effect of Heating Mode and Copper Content on
the Densification of W-Cu Alloys

Avijit Mondal,1 Anish Upadhyaya,2 and Dinesh Agrawal3


1
NTPC Energy Technology Research Alliance (NETRA), NTPC Ltd., E-III, Echotech-2, Greater Noida 201308, India
2
Department of Materials Science and Engineering, Indian Institute of Technology, Kanpur 208016, India
3
Materials Research Institute, The Pennsylvania State University, University Park, PA 16802, USA

Correspondence should be addressed to Avijit Mondal; avijit.mondal@gmail.com

Received 18 June 2013; Accepted 13 July 2013

Academic Editors: D. Das and Y. Ge

Copyright © 2013 Avijit Mondal et al. This is an open access article distributed under the Creative Commons Attribution License,
which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

This study investigates the effect of heating mode on the sintering of tungsten-copper alloys containing up to 30 wt.% Cu. The
sinterability of the W-Cu system consolidated in a 2.45 GHz multimode microwave furnace has been critically compared with that
processed in a radiatively heated (conventional) furnace. The as-pressed W-Cu alloys can be readily sintered in microwave furnace
with substantial (sixfold) reduction in the processing time. As compared to conventional sintering, microwave processing results
in greater densification, more homogenous distribution of the binder phase, and smaller tungsten grain size. The densification in
compacts increases with increasing Cu content. For all compositions, the electrical conductivity and hardness of microwave sintered
W-Cu alloys are higher than those of their conventionally sintered counterparts. This study investigates the effect of heating mode
on the sintering of tungsten-copper alloys containing up to 30 wt.% Cu. The W-Cu alloys were sintered in a 2.45 GHz microwave
furnace with substantial (sixfold) reduction in the processing time. As compared to conventional sintering, microwave processing
results in greater densification, more homogenous distribution of the binder phase, and smaller tungsten grain size. This results in
higher electrical conductivity and hardness of the microwave sintered W-Cu alloys.

1. Introduction enhance densification during sintering [7]. However, such a


technique is not economically viable for upscaling. Another
Tungsten-copper alloys are widely used for a range of approach for ensuring densification in the W-Cu system
applications, such as electrical contact materials, thermal- is through the addition of transition metal additives that
management devices, and in ordnance applications [1–3]. activate the sintering kinetics [8–10]. Such activators though
On account of the refractoriness of the major phase (Wm.p. : remain segregated at the grain boundaries and thereby result
3420∘ C), these alloys are usually processed through liquid in poor mechanical properties [11]. Consequently, most of the
phase sintering [4]. Due to lack of solubility between W efforts are in consolidation of W-Cu alloys through the use of
and Cu, it is rather difficult to attain full densification in ultrafine constituent powders [12]. Such powders are prone
this system. For ensuring complete densification through to contamination and therefore require careful handling.
liquid phase sintering, sufficient amount of melt and its Most importantly, when processed through a convention-
homogeneous distribution are required [5]. Unfortunately, ally adopted technique, the compacts from these powders
due to the density difference between the constituents, it undergo significant microstructural coarsening. It is there-
is rather difficult to ensure Cu homogeneity. Thus, the W fore well recognized that a faster sintering cycle would restrict
and Cu powders are often comilled, which, in turn, results microstructural coarsening. However, because sintering is
in contamination from the milling media and results in a diffusion-driven process, the effect of high heating rate
subsequent degradation of property, such as conductivity [6]. on the densification response is not that straightforward
Alternatively, Cu coating over the tungsten has shown to [13]. Furthermore, most conventional sintering furnaces are
2 Indian Journal of Materials Science

resistance heated and the heat transfer is limited by radiation 1400


and conduction. Hence, to ensure uniformity of temperature Microwave Conventional
distribution, the samples are usually heated at a slower rate 1200
and for longer times.
1000
In the recent years, on account of the interaction of

Temperature (∘ C)
microwaves with the particulate material, microwave sinter-
800
ing has been widely used for consolidating a range of materi-
als [14]. Till very recently, the application of microwaves was 600
restricted to ceramics and cermets [15]. However, Roy et al.
[16] demonstrated that particulate metals too can effectively 400
interact with microwaves and get heated up. This has opened
the possibility for consolidating a range of particulate metals 200
and their alloys through microwaves sintering [17]. So far, for
the metallic systems, microwave sintering has been applied 0
either for pure metals or for multicomponents systems which 0 100 200 300 400 500
have mutual interaction, such as Fe-Cu, Cu-Sn, and W- Time (min)
Ni-Fe alloys [18–24]. This study evaluates the feasibility of W-10Cu
microwave sintering of W-Cu system where the constituents W-20Cu
are chemically noninteractive. W-30Cu

Figure 1: Comparison of heating profile of conventionally and


microwave sintered W-Cu alloys.
2. Experimental Section
For the current study, the sintering of W-Cu particulate
compacts having 10, 20, and 30 wt.% copper was investigated. was also expressed in terms of a normalized, dimensionless
The starting powder in as-mixed condition was supplied by densification parameter which is expressed as follows:
NEI Corporation, Somerset, NJ, USA. The tungsten powder
size was 70 nm. The as-received powders were uniaxially densification parameter
compacted in a 50 T hydraulic press (model: CTM-50,
supplier: FIE, Ichalkaranji, India) equipped with a floating (sintered density − green density) (1)
= .
die. To facilitate compaction, zinc stearate was used as die- (theoretical density − green density)
wall lubricant. Cylindrical compacts having 16 mm diameter
and 8 mm height were pressed at 200 MPa. To take into The samples of sintered W-Cu alloys were prepared for
account the variability of the density of the compacts due microstructural investigation using standard metallographic
to the variation in the composition, the as-pressed (green) practices. The samples were wet-polished in a manual pol-
and the sintered densities were normalized with respect isher (model: Lunn Major, supplier: Struers, Denmark) using
to the theoretical density which was calculated using the a series of 6 𝜇m, 3 𝜇m, and 1 𝜇m diamond pastes, which
inverse rule of mixture. The green density for all the three was followed by cloth-polishing using an 0.04 𝜇m colloidal
compositions was nearly the same and varied 50 to 53% with silica suspension. The microstructural observations were
the theoretical density. done using a scanning electron microscope (model: Quanta,
The as-pressed compacts were sintered using both a supplier: FEI, The Netherlands) both in secondary as well as
radiatively heated (conventional) as well as a 2.45 GHz back-scattered imaging modes.
multimode microwave furnaces. The former was carried in a The bulk hardness of the conventional and microwave
MoSi2 -heated horizontal tubular furnace (model: OKAY 70T, sintered W-Cu alloys was assessed using a Vickers hardness
supplier: Bysakh Inc., Kolkata, India) at a constant heating tester (model: V100-C1, supplier: Leco, Japan) under 50 N
rate of 5∘ C/min. To ensure uniform temperature distribution load. To take into account the variability, for each sample, on
during heating, intermittent isothermal hold for 30 min was an average, ten hardness indentations were taken randomly
provided at 500∘ C and 900∘ C. For microwave sintering, a across the cross-section. The electrical conductivity of the
6 kW microwave furnace (model: RC/20SE, supplier: Amana sintered W-Cu alloys was evaluated using a digital conduc-
Radarange) was employed. Further details of the microwave tivity meter (model: 757, supplier: TechnoFour, Pune, India).
sintering unit are provided elsewhere [20, 21]. All W-Cu
compacts were sintered at 1200∘ C in hydrogen atmosphere 3. Results and Discussion
with dew point −35∘ C. At the sintering temperature, the
compacts were kept isothermally for 30 min. Figure 1 compares the thermal profiles of the W-Cu compacts
The density of the sintered compacts was evaluated both consolidated in a conventional and microwave furnace. It is
through dimensional measurements as well as Archimedes’ interesting to note that all compacts (W-10Cu, W-20Cu, and
density measurements. To take into account the effect of the W-30Cu) exhibit very strong coupling with microwaves and
initial as-pressed density variation, the compact sinterability undergo very rapid heating. As compared to conventional
Indian Journal of Materials Science 3

100
Sintered density (%theoretical) 1.0

90 0.9

Densification parameter
80 0.8

70 0.7

60 0.6

50 0.5
W-10Cu W-20Cu W-30Cu W-10Cu W-20Cu W-30Cu
Cu content (wt.%) Cu content (wt.%)

Conventional Conventional
Microwave Microwave
(a) (b)

Figure 2: Effect of heating mode on the (a) sintered density and (b) densification parameter of W-Cu alloys containing varying amounts of
Cu. All compacts were sintered at 1200∘ C.

sintering, microwave heating results in about more than six- similar interaction of microwaves with tungsten-based alloys
fold decrease in the heating time. For conventional sintering, [24, 27–29].
to ensure uniform homogenous heat transfer during heating Figure 2(a) compares the effect of Cu content and the
and prevent thermal shock to the furnace assembly (tube heating mode on the sintered density of the W-Cu compacts.
and heating elements), the heating rate was restricted to The corresponding densification parameters are summarized
5∘ C/min. In comparison, because of the direct coupling of the in Figure 2(b). From Figure 2, it is interesting to note that the
microwaves with the powder, an overall high heating rate of densification of the W-Cu alloys is enhanced with increasing
about 35∘ C/min was achieved. From Figure 1, it is evident that copper content. It can also be inferred that, irrespective of the
W-Cu compacts can be consolidated in microwave furnace copper content, the density of microwave sintered compacts
in significantly (∼85%) lower time that those achieved in a is higher than that achieved through conventional sintering.
conventional furnace. Mishra et al. [25] have modeled the This underscores the efficacy of microwaves processing not
microwave heating of metals and have elucidated its mech- only from the viewpoint of sintering time compression but
also in densification enhancement standpoint as well. The
anism in metallic system as mainly joule heating. Usually, for
role of increasing copper content on the densification can
conductive materials, such as metals, the interaction with the
be attributed to the greater capillary stresses induced grain
microwaves is restricted at the surface and results in eddy rearrangement [30]. During conventional sintering, however,
currents. This depth of penetration in metals, also known as prior to Cu melt formation, substantial solid-state densifica-
skin-depth (𝛿), is defined as the distance into the material tion also occurs which results in W-W particle bonding which
at which the incident power drops to 1/e (36.8%) of the results in formation of a skeletal structure [31, 32]. In most
surface value. This skin depth is indirectly proportional to the multiphase systems, the occurrence of such structure is a
conductivity of the metals and increases with temperatures. common occurrence [33, 34]. However, upon melt formation,
Usually, for most common metallic systems, the skin depth the liquid chemically dissolves the interparticle bonds and
varies in the range of 0.5 to 20 𝜇m. By corollary, therefore, disintegrates the skeletal structure. In W-Cu system, owing
microwaves will interact volumetrically if the metallic system to the insolubility of W in copper [35], the Cu melt does
is in particulate form and contains porosity [26]. For W not chemically dissolve the W-W interparticle neck formed
and Cu, the skin depth varies between 2.5–6 𝜇m and 1.75– during solid-state sintering. In addition to poor solubility, the
2.25 𝜇m, respectively. For tungsten, the initial particle sizes W-Cu has a high dihedral angle (∼90∘ ) [30, 36]. The dihedral
are much smaller as compared to the skin depth, hence angle can be expressed as a balance of the interfacial W-Cu
it is reasonable to assume that all the tungsten powders melt and W-W energy as well as geometrically in terms of
get homogenously and volumetrically heated up. The W- the interparticle W-W neck size (𝑋) to the W-grain size (𝐺W )
Cu powder due to its poor compressibility and had up to ratio as follows [37]:
50% porosity in the green state. This coupling with the small
𝛾 𝑋
powder size ensures good interaction of the compact with the 𝜙 = 2cos−1 ( WW ) = 2sin−1 ( ). (2)
microwaves. Elsewhere, other researchers too have reported 𝛾W-Cu 𝐺W
4 Indian Journal of Materials Science

SEM micrograph

Conventional Microwave
(a)
SEM micrograph

Conventional Microwave
(b)
SEM micrograph

Conventional Microwave
(c)

Figure 3: SEM photomicrographs of (a) W-10Cu, (b) W-20Cu, and (c) W-30Cu alloys consolidated in a radiatively heated (conventional)
and microwave furnaces.

In the W-Cu system, the W-W interfacial energy is rela- densification due to capillary-induced rearrangement. To
tively low (2.79 J/m2 ) [38]; this entails that the interparticle circumvent this problem, it is essential that the W-W bond
tungsten bonds that form through solid-state sintering are formation prior to melting is restricted. The rapid heating in
energetically favoured and therefore stable. Thus, on account microwaves helps in achieving this. Thus, it is not surprising
of both poor intersolubility as well as high dihedral angles, that, as compared to conventional heating, microwave
the W-Cu system is fundamentally not predisposed for sintering results in higher densification. Elsewhere also,
Indian Journal of Materials Science 5

SEM micrograph

W grains Cu pool

Conventional Microwave
(a) (b)

Figure 4: Microstructure of W-30Cu alloys sintered at 1200∘ C in (a) conventional and (b) microwave furnaces.

Table 1: Effect of Cu content and heating mode on the grain size, bulk hardness, and electrical conductivity of sintered W-Cu alloys.

Composition (in wt.%) Heating mode∗ Grain Size (nm) Electrical Conductivity (% IACS) Hardness (HV5 )
CON 511 ± 91 16.0 ± 0.5 224 ± 19
W-10Cu
MW 362 ± 104 21.2 ± 0.3 248 ± 12
CON 496 ± 108 32.6 ± 0.2 280 ± 16
W-20Cu
MW 403 ± 105 35.2 ± 0.5 349 ± 19
CON 499 ± 75 40.0 ± 0.3 258 ± 8
W-30Cu
MW 302 ± 63 43.1 ± 0.7 322 ± 9

CON: conventional; MW: microwave.

researchers [13, 38, 39] have proposed that slow heating rates Upadhyaya and German [30] and Mondal et al. [29, 41] on liq-
are detrimental to rearrangement. However, in conventional uid phase sintered W-Ni alloys containing Fe and Cu. How-
resistance heated furnace where the heat transfer mode ever, unlike W-Ni-Fe and W-Ni-Cu systems, due to lack of
is predominantly radiative and convective, high heating W solubility in Cu, solution reprecipitation is not expected to
rates are not readily and economically feasible. Microwave contribute to the tungsten grain coarsening [42]. For the W-
consolidation therefore offers a viable alternative to process Cu system, the grain growth primarily occurs through coa-
W-Cu alloys through liquid phase sintering. lescence for which W-W particle contact is a prerequisite. A
Figures 3(a) to 3(c) compare distribution of the binder higher Cu volume fraction in the alloys reduces the tungsten
(Cu) phase in the microstructures of conventionally and contiguity in the structure and thereby results in less coarsen-
microwave sintered W-Cu alloys with varying Cu content. ing [43]. In case of microwave sintering, besides the less time
Figures 4(a) and 4(b) compare microstructures (at higher
available for microstructural coarsening, a more uniform
magnification) of liquid phase sintered W-30Cu alloys con-
melt distribution also contributes towards restricting the
solidated in a conventional and microwave furnace, respec-
grain growth by further reducing the interparticle contacts.
tively. From Figures 3 and 4, it is evident, that in con-
ventional sintering, copper is more inhomogeneously dis- Table 1 also compares the effect of heating mode and the
tributed. Moreover, there is evidence of binder pools which Cu content on the tungsten grain size, electrical conductivity,
is indicative of poor rearrangement of the bonded skeletal and the hardness. As expected, a Cu content in general results
structure that results in densification during liquid phase in an increase in the electrical conductivity. However, it has
sintering occurring through pore-filling effect [40]. In a very an opposing effect on the compact hardness. It is rather
recent report, Luo et al. [28] also made similar observations. interesting to note that, as compared to conventional sinter-
However, their assessment was qualitative and the reason ing, the microwave sintered compacts have higher electrical
for this was not clearly elaborated. Table 1 summarizes the conductivity and greater hardness. This can be correlated to
W grain sizes measured in the sintered compacts. For both a sintered density, more uniform binder distribution, and
sintering modes, the tungsten grain size reduces as the copper a smaller tungsten grain size achieved in the W-Cu alloys
content increases. Similar observations were also reported by consolidated in microwave furnace.
6 Indian Journal of Materials Science

4. Conclusions [10] V. Gauthier, F. Robaut, A. Upadhyaya, and C. H. Allibert, “Effect


of Fe on the constitution of Cu-W alloys at 1200∘ C,” Journal of
In summary, this study shows that the W-Cu alloys with Alloys and Compounds, vol. 361, no. 1-2, pp. 222–226, 2003.
varying Cu content can be successfully consolidated through [11] J. L. Johnson and R. M. German, “Factors affecting the ther-
microwave sintering. As compared to conventional sintering, mal conductivity of W-Cu composites,” in Proceedings of the
the W-Cu alloys were consolidated in microwaves with about International Conference & Exhibition on Powder Metallurgy &
85% reduction in the overall processing time. Irrespective Particulate Materials, vol. 4 of Advances in Powder Metallurgy
of the Cu content, microwave sintered compacts had higher and Particulate Materials, pp. 201–213, Metal Powder Industries
densification and a more uniform microstructure with lower Federation, May 1993.
tungsten grain size. Consequently, the electrical conductiv- [12] J. L. Johnson, J. J. Brezovsky, and R. M. German, “Effects
ity and the bulk hardness of W-Cu sintered alloys using of tungsten particle size and copper content on densification
microwaves were higher than those of their counterparts of liquid-phase-sintered W-Cu,” Metallurgical and Materials
Transactions A, vol. 36, no. 10, pp. 2807–2814, 2005.
processed through conventional (radiatively heated) furnace.
[13] R. M. German and S. Farooq, “An update on the theory of liquid
phase sintering,” in Sintering’87, S. Sōmiya, M. Shimada, M.
Acknowledgments Yoshimura, and R. Watanabe, Eds., vol. 1, pp. 459–464, Elsevier
Science, New York, NY, USA, 1988.
The authors thank NEI Corporation, Somerset, NJ, USA, [14] M. Gupta and W. L. E. Wong, Microwaves and Metals, John
for supplying the powder used for this investigation. The Wiley & Sons, Singapore, 2007.
financial support from the Indo-US Science and Technology [15] K. Rödiger, K. Dreyer, T. Gerdes, and M. Willert-Porada,
Forum (IUSSTF), New Delhi, for this study is gratefully “Microwave sintering of hardmetals,” International Journal of
acknowledged. Refractory Metals and Hard Materials, vol. 16, no. 4–6, pp. 409–
416, 1998.
[16] R. Roy, D. Agrawal, J. Cheng, and S. Gedevanishvili, “Full
References sintering of powdered-metal bodies in a microwave field,”
Nature, vol. 399, pp. 668–670, 1999.
[1] A. Upadhyaya, “Processing strategy for consolidating tungsten
heavy alloys for ordnance applications,” Materials Chemistry [17] K. Saitou, “Microwave sintering of iron, cobalt, nickel, copper
and Physics, vol. 67, no. 1–3, pp. 101–110, 2001. and stainless steel powders,” Scripta Materialia, vol. 54, no. 5,
pp. 875–879, 2006.
[2] A. Upadhyaya, “Processing strategy for consolidating tungsten
heavy alloys for ordnance and thermal-management applica- [18] R. M. Anklekar, D. K. Agrawal, and R. Roy, “Microwave sin-
tions,” Transactions of the Indian Institute of Metals, vol. 55, no. tering and mechanical properties of PM copper steel,” Powder
1-2, pp. 51–69, 2002. Metallurgy, vol. 44, no. 4, pp. 355–362, 2001.
[19] M. Gupta and W. L. E. Wong, “Enhancing overall mechan-
[3] R. M. German, K. F. Hens, and J. L. Johnson, “Powder
ical performance of metallic materials using two-directional
metallurgy processing of thermal management materials for
microwave assisted rapid sintering,” Scripta Materialia, vol. 52,
microelectronic applications,” International Journal of Powder
no. 6, pp. 479–483, 2005.
Metallurgy, vol. 30, no. 2, pp. 205–215, 1994.
[20] S. S. Panda, V. Singh, A. Upadhyaya, and D. Agrawal, “Sintering
[4] A. Belhadjhamida and R. M. German, “Tungsten and tungsten response of austenitic (316L) and ferritic (434L) stainless steel
alloys by powder metallurgy: a status review,” in Tungsten and consolidated in conventional and microwave furnaces,” Scripta
Tungsten Alloys Recent Advances, A. Crowson and E. S. Chen, Materialia, vol. 54, no. 12, pp. 2179–2183, 2006.
Eds., pp. 3–20, TMS, Warrendale, Pa, USA, 1991.
[21] S. S. Panda, V. Singh, A. Upadhyaya, and D. Agrawal, “Effect
[5] J. L. Johnson, J. J. Brezovsky, and R. M. German, “Effect of of conventional and microwave sintering on the properties
liquid content on distortion and rearrangement densification of yttria alumina garnet-dispersed austenitic stainless steel,”
of liquid-phase-sintered W-Cu,” Metallurgical and Materials Metallurgical and Materials Transactions A, vol. 37, no. 7, pp.
Transactions A, vol. 36, no. 6, pp. 1557–1565, 2005. 2253–2264, 2006.
[6] J.-C. Kim, S.-S. Ryu, Y. D. Kim, and I.-H. Moon, “Densification [22] C. Padmavathi, A. Upadhyaya, and D. Agrawal, “Corrosion
behavior of mechanically alloyed W-Cu composite powders by behavior of microwave-sintered austenitic stainless steel com-
the double rearrangement process,” Scripta Materialia, vol. 39, posites,” Scripta Materialia, vol. 57, no. 7, pp. 651–654, 2007.
no. 6, pp. 669–676, 1998. [23] A. Upadhyaya and G. Sethi, “Effect of heating mode on the
[7] A. Upadhyaya and C. Ghosh, “Effect of coating and activators densification and microstructural homogenization response of
on sintering of W-Cu alloys,” Powder Metallurgy Progress, vol. premixed bronze,” Scripta Materialia, vol. 56, no. 6, pp. 469–472,
2, no. 2, pp. 98–110, 2002. 2007.
[8] J. L. Johnson and R. M. German, “A theory of activated liquid [24] A. Upadhyaya, S. K. Tiwari, and P. Mishra, “Microwave sinter-
phase sintering and its application to the W-Cu system,” in ing of W-Ni-Fe alloy,” Scripta Materialia, vol. 56, no. 1, pp. 5–8,
Proceedings of the International Conference & Exhibition on 2007.
Powder Metallurgy & Particulate Materials, vol. 3 of Advances in [25] P. Mishra, G. Sethi, and A. Upadhyaya, “Modeling of microwave
Powder Metallurgy and Particulate Materials, pp. 35–46, Metal heating of particulate metals,” Metallurgical and Materials
Powder Industries Federation, June 1992. Transactions B, vol. 37, no. 5, pp. 839–845, 2006.
[9] D.-G. Kim, G.-S. Kim, M.-J. Suk, S.-T. Oh, and Y. D. Kim, “Effect [26] A. Mondal, D. Agrawal, and A. Upadhyaya, “Microwave heating
of heating rate on microstructural homogeneity of sintered of pure copper powder with varying particle size and porosity,”
W-15 wt% Cu nanocomposite fabricated from W-CuO powder The Journal of Microwave Power and Electromagnetic Energy,
mixture,” Scripta Materialia, vol. 51, no. 7, pp. 677–681, 2004. vol. 43, no. 1, pp. 5–10, 2009.
Indian Journal of Materials Science 7

[27] G. Prabhu, A. Chakraborty, and B. Sarma, “Microwave sintering


of tungsten,” International Journal of Refractory Metals and Hard
Materials, vol. 27, no. 3, pp. 545–548, 2009.
[28] S. D. Luo, J. H. Yi, Y. L. Guo, Y. D. Peng, L. Y. Li, and J.
M. Ran, “Microwave sintering W-Cu composites: analyses of
densification and microstructural homogenization,” Journal of
Alloys and Compounds, vol. 473, no. 1-2, pp. L5–L9, 2009.
[29] A. Mondal, A. Upadhyaya, and D. Agrawal, “Microwave sin-
tering of refractory metals/alloys: W, Mo, Re, W-Cu, W-Ni-
Cu and W-Ni-Fe alloys,” The Journal of Microwave Power and
Electromagnetic Energy, vol. 44, no. 1, pp. 28–44, 2010.
[30] A. Upadhyaya and R. M. German, “Densification and dilation
of sintered W-CU alloys,” International Journal of Powder
Metallurgy, vol. 34, no. 2, pp. 43–55, 1998.
[31] J. L. Johnson and R. M. German, “Solid-state contributions to
densification during liquid-phase sintering,” Metallurgical and
Materials Transactions B, vol. 27, no. 6, pp. 901–909, 1996.
[32] V. V. Panichkina and N. I. Filipov, “Structure formation pro-
cesses in the sintering of disperse copper-tungsten pseudoal-
loys,” Science of Sintering, vol. 26, pp. 269–275, 1994.
[33] G. S. Upadhyaya, Sintered Metallic and Ceramic Materials, John
Wiley & Sons, New York, NY, USA, 2000.
[34] A. P. Savitskii, Liquid Phase Sintering of the Systems with
Interacting Components, Russian Academy of Sciences, Tomsk,
Russia, 1993.
[35] H. Okamoto, P. R. Subramaniam, and L. Kacprzak, Binary Phase
Diagrams, vol. 2, American Society for Microbiology, Materials
Park, Ohio, USA, 2nd edition, 1990.
[36] A. Upadhyaya and R. M. German, “Shape distortion in liquid-
phase-sintered tungsten heavy alloys,” Metallurgical and Mate-
rials Transactions A, vol. 29, no. 10, pp. 2631–2638, 1998.
[37] R. M. German, P. Suri, and S. J. Park, “Review: liquid phase
sintering,” Journal of Materials Science, vol. 44, no. 1, pp. 1–39,
2009.
[38] B. D. Storozh and P. S. Kislyi, “Sintering of tungsten-alumina
cermets in the presence of a liquid phase,” Soviet Powder
Metallurgy and Metal Ceramics, vol. 13, no. 9, pp. 712–716, 1974.
[39] R. Bollina and R. M. German, “Heating rate effects on
microstructural properties of liquid phase sintered tungsten
heavy alloys,” International Journal of Refractory Metals and
Hard Materials, vol. 22, no. 2-3, pp. 117–127, 2004.
[40] S.-M. Lee and S.-J. L. Kang, “Theoretical analysis of liquid-phase
sintering: pore filling theory,” Acta Materialia, vol. 46, no. 9, pp.
3191–3202, 1998.
[41] A. Mondal, A. Upadhyaya, and D. Agrawal, “Microwave sinter-
ing of W-18Cu and W-7Ni-3Cu alloys,” The Journal of Microwave
Power and Electromagnetic Energy, vol. 43, no. 1, pp. 11–16, 2009.
[42] W. J. Huppmann, “The elementary mechanisms of liquid phase
sintering: II. solution-reprecipitation,” Zeitschrift fuer Metal-
lkunde, vol. 70, pp. 792–797, 1979.
[43] S.-C. Yang, S. S. Mani, and R. M. German, “The effect of
contiguity on growth kinetics in liquid-phase sintering,” Journal
of Metals, vol. 42, no. 4, pp. 16–19, 1990.
BioMed Research Smart Materials
International Research

International Journal of
Corrosion
Journal of Journal of
Nanotechnology
Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation
Composites
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013

International Journal of Journal of


Polymer Science Metallurgy

Hindawi Publishing Corporation Hindawi Publishing Corporation


http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013

Nanomaterials
Submit your manuscripts at
http://www.hindawi.com

Journal of
Materials Advances in
Materials Science and Engineering
Hindawi Publishing Corporation Volume 2013 Hindawi Publishing Corporation
http://www.hindawi.com http://www.hindawi.com Volume 2013

Nanomaterials
Journal of

The Scientific
Scientifica
International Journal of
Journal of
Nanoparticles
Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation
World Journal
Hindawi Publishing Corporation
Biomaterials
Hindawi Publishing Corporation
http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013

ISRN ISRN ISRN ISRN ISRN


Nanotechnology Polymer Science Materials Science Corrosion Ceramics
Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation Hindawi Publishing Corporation
http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013 http://www.hindawi.com Volume 2013

You might also like