You are on page 1of 8

Int.

Journal of Refractory Metals & Hard Materials 28 (2010) 550–557

Contents lists available at ScienceDirect

Int. Journal of Refractory Metals & Hard Materials


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

Densification behaviour of pure molybdenum powder by spark plasma sintering


R. Ohser-Wiedemann *, U. Martin, H.J. Seifert, A. Müller
Institute of Materials Science, TU Bergakademie Freiberg, Gustav-Zeuner-Str. 5, 09599 Freiberg, Germany

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

Article history: Pure molybdenum was sintered with SPS under various temperatures, external pressures and heating
Received 18 December 2009 rates. The microstructure of the specimens representing the different sintering conditions was investi-
Accepted 15 March 2010 gated by classical metallographic methods. The relative density, the microhardness and the chord length
distribution were measured. Linear shrinkage, depending on time or temperature, was calculated from
piston travel, which was recorded during sintering process. These results show that the main part of con-
Keywords: solidation takes place during fast heating up. The densification behaviour is controlled mainly by sinter-
Molybdenum
ing temperatures and applied pressure. The molybdenum powder was successfully consolidated by SPS in
Spark plasma sintering
Densification
very short times. A relative density of 95% was reached by sintering temperatures of 1600 °C and external
Microstructure pressure of 67 MPa.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction This allows the sintering of nanosized or metastable powders


close to the theoretical density with slight grain growth or reten-
Pure molybdenum (Mo) is an important refractory metal used tion of metastability as well as cleaned grain boundaries. Further
for a wide scale of engineering applications, due to advantageous advantages are sintering of powders without any additives, no
properties. Some examples are ribbons and wires for lighting tech- need of cold compaction and less sensitivity to initial powders
nology, semiconductor base plates for power electronics, elec- characteristics. SPS is therefore an economical alternative to con-
trodes for glass melting, parts for high-temperature furnaces, ventional sintering [1–9]. However, the densification mechanism
spraying wires for automotive industry and sputter targets or during SPS process is unclear as yet [8]. It is assumed that small
evaporation sources for coating technology. Components from spark discharges appear between individual powder particles caus-
pure molybdenum or its alloys are produced either by powder ing locally high temperatures in few minutes. These high temper-
metallurgy or by melting processes. Generally, powder metallurgi- atures induce vaporization or melting of the powder surface,
cal route is preferred since a fine-grained microstructure is ob- destruction of oxide layers and raising sinter activity. Necks be-
tained, which improve the mechanical properties of final tween the powder particles are formed. This and the high compres-
products considerably. Additionally, the powder route is the exclu- sive force lead to the sintering of the particles [2,3,5,10]. However,
sive possibility for production of molybdenum–copper–alloys or the existence of spark plasma in SPS process is not clearly demon-
for doping molybdenum with high-temperature resistant oxides strated, in particular when non-conductive powders are used.
[1]. However, the fundamental effect of electric current on mass trans-
Spark plasma sintering (SPS) is a new technology for compac- port has been clearly shown, elucidated in [8]. The electrical cur-
tion of metallic or ceramic powders. It is a short-time sintering rent affects also the diffusion kinetics of reaction between
process, where powder particles are compacted by uniaxial press- different phases, for example between Mo and Si, see [6]. Early
ing and heating simultaneously. The heating results from a pulsed densification stages of nanocrystalline ceramics by SPS were dis-
electric field and a high heating rate of more than 100 K/min can be cussed by Chaim [7], where the author identified two atomistic
achieved. mechanisms. Nanocrystalline ceramics with low yield stress were
Compared to conventional sintering methods applying external mainly compacted by plastic deformation. Ceramics with high
pressure like Hot Pressing (HP) or Hot-Isostatic-Pressing (HIP), yield stress consolidate dominantly by collective grain rotation
densification by SPS is extremely fast. Thus, the sintering temper- and sliding. This is a hint on the important role of applied pressure
atures can be lower which limits the grain growth. during spark plasma sintering.
First studies on short-time sintering of molybdenum powders
by Plasma Pressure Compaction (P2C) are reported in [9,11,12].
Similar to the SPS process, the powders were filled in a graphite
* Corresponding author. Tel.: +49 3731 39 2647; fax: +49 3731 39 3657.
E-mail address: ohser@ww.tu-freiberg.de (R. Ohser-Wiedemann). die. The heating by P2C was carried out at first with pulsed electric

0263-4368/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijrmhm.2010.03.003
R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557 551

current up to 1100 °C, followed by heating with constant current


up to the final sintering temperature. Microsized powders (average
size 47 lm) were consolidated at 1650 °C, 48 MPa, for 1–2 min, up
to a relative density of 98%. ‘‘Nanosized” powders (average size
0.1 lm) were sintered at lower temperatures (1400 °C, 48 MPa,
3 min) up to a relative density of 97%. The smaller grain size of
nanocrystalline samples leads to a higher microhardness (2.95
GPa) in comparison to the microcrystalline samples (2.16 GPa).
Clean grain boundaries coupled with uniform grain structure were
taken as evidence for the occurrence of surface cleaning during
plasma activation stage.
Furthermore, Mo–Si–B multiphase alloys were successfully con-
solidated by SPS at 1200 °C and an applied pressure of 40 MPa [13].
Unfortunately, the relative densities after sintering are not
specified.
Fig. 2. Typical SPS runs with different heating rates.
The motivation of the present work was to estimate the best SPS
conditions for consolidation of pure molybdenum powder and to
densities were calculated based on the theoretical density of
achieve information about the densification behaviour of molybde-
10.28 g cm 3, listed in [1].
num during this process, compared with short-time sintering by
The microstructure was studied by optical microscopy as well
P2C and isothermal sintering.
as by scanning electron microscopy at cross sections (sections par-
allel to the acting force). The preparation steps were grinding,
mechanical polishing up to 3 lm diamond abrasive, followed by
2. Experimental short electrolytic polishing and chemical etching using Murakami’s
reagent (a solution mixture of 2.5 g potassium hydroxide, 2.5 g
Pure molybdenum powders (99.95 wt%) with particle size of 3– potassium ferricyanide and 50 ml of distilled water).
5 lm (produced by Plansee Metall GmbH Austria) were consoli- For the batches 3.1–3.7, the grain size was estimated by con-
dated by SPS. Before consolidation, the powder contains an oxygen ventional metallographic methods. In our setting, the grain size
impurity of 400 ppm, as indicated by the producer of the material. distribution is defined as the chord length distribution, which
The powder particles (Fig. 1) are of spherical shape, some of them was measured at least as five measuring fields per cross sections
include small pores. The particles agglomerates and tend to form using the software ‘‘a4i Analysis” of aquinto AG.
chains. The powders contain a fraction of fine particles with sizes The hardness of bulk molybdenum specimens was measured at
<1 lm. cross sections either by nanohardness tester (CSM Instruments) or
The powders were filled without any additives in cylindrical by microhardness tester (LECO Instrumente GmbH). The nanoin-
graphite dies either of 20 or 40 mm in diameter. The sintering runs dentations were generated using a BERKOVICH-indenter, which
in a FCT-HP D 25 spark plasma sinter equipment, manufactured by was loaded with a maximum load of 30 mN. The indentations
FCT Systeme GmbH (Germany). The compaction of all batches (20 in samples with high density, 40 in samples with low density)
starts with raising the external pressure up to 29, 57 and 67 MPa, were individually located inside of grains, away from grain bound-
respectively, followed by heating up to temperatures between aries and pores. Basing on the recorded load-displacement-curve,
850 and 2000 °C under vacuum. The heating rate was diversified the indentation hardness and the indentation modulus were calcu-
between 130 and 360 K/min, and the holding time at maximum lated as described in [16]. The microhardness HV0.1 was measured
temperature was 3 min for all batches. The pulse pattern of the with a VICKERS indenter.
electric current was adjusted on 2:1 (10 ms pulse time, 5 ms pause The X-ray diffraction (XRD) spectra of the carburized outer zone
time). The compaction pressure decreased during cooling step. were recorded with URD65 (Seifert FPM) powder diffractometer
Fig. 2 gives an example of two typical SPS runs. Table 1 contains using Cu–Ka radiation and a graphite monochromator in the dif-
sintering parameters of all batches. fracted beam. The PDF data base [17] was used for qualitative
The densities of all consolidated bulk molybdenum specimens phase analysis. Quantitative pole figures were obtained by measur-
were measured by the Archimedes technique [14,15]. The relative ing with the XRD system D8-Discover from Bruker AXS using Co–
Ka radiation.

3. Experimental results

3.1. Density

The measured densities and the calculated relative densities of


all sintered samples are listed in Table 1. Fig. 3 shows the relative
densities depending on sintering temperatures and pressures. The
correlation between relative density and temperature follows the
well-known sigmoid course of heating up during isothermal sinter-
ing. A relative density of 95% and higher can be reached if the sin-
tering temperature excites 1600 °C and the external pressure
amounts 57 MPa or more. In the experiments with low external
pressure of 29 MPa, the relative density does not exceed 95% even
at high temperatures. This indicates the applied pressure as an
important factor for the densification during the SPS process. The
Fig. 1. Molybdenum powder before consolidation. heating rates, studied in the present article, do not influence the
552 R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557

Table 1
Sintering parameters and results of density and nanohardness measurements.

Batch Heating rate (K/ Final sintering Pressure Final density (g/ Final relative Indentation hardness Indentation modulus
min) temperature (°C) (MPa) cm3) density (%) (MPa) (GPa)
1.1 130 1600 29 9.19 89.4 1766 ± 194 214 ± 22
1.2 130 1800 29 9.47 92.1 1646 ± 185 280 ± 43
1.3 130 1900 29 9.50 92.4 1810 ± 138 227 ± 21
2.1 200 1400 57 9.30 90.5 1765 ± 194 214 ± 22
2.2 130 1600 57 9.70 94.4 2009 ± 169 246 ± 16
2.3 360 1800 57 9.80 95.3 1934 ± 88 281 ± 30
2.4 360 1900 57 9.80 95.3 2025 ± 143 288 ± 26
3.1 200 850 67 5.96 58.0 – –
3.2 325 1100 67 7.18 69.0 1504 ± 279 123 ± 13
3.3 200 1400 67 9.24 89.9 1665 ± 220 233 ± 28
3.4 280 1600 67 9.77 95.0 2085 ± 99 260 ± 21
3.5 340 1800 67 9.84 95.7 2152 ± 96 289 ± 29
3.6 360 1900 67 9.85 95.8 2151 ± 84 291 ± 31
3.7 200 2000 67 9.88 96.1 1751 ± 91 288 ± 31

Fig. 3. Relative density of spark plasma sintered molybdenum vs. sintering


temperature, compared with other published data.
Fig. 4. Results of nanohardness measurements depending on sintering temperature
and relative density.

density of molybdenum samples, significantly. This is in accor-


dance with the densification behaviour of copper powders with
and indentation modulus (Fig. 4, Table 1). Sintered bodies with
SPS [10]. Finally, it is observed that the density of spark plasma sin-
low density apparently exhibit lower hardness and modulus.
tered bulk molybdenum is primarily controlled by external pres-
If pores occur beneath grains, the indentation process is over-
sure and temperature.
lapped by elastic recovery, plastic deformation or crack propaga-
Our results are in good agreement with the P2C results in
tion. In order to exclude the influence of porosity, measured
[9,11,12], where a relative density of 98% was measured at
hardness values of each sample were divided by their relative den-
1650 °C and at 48 MPa external pressure. Additionally, we compare
sity. In this case, the indentation hardness’s of all investigated sam-
the densities obtained by SPS with those of isothermal sintering.
ples are in the same range, independent on sintering conditions.
The isothermal powder metallurgical route starts with hydraulic
The average indentation hardness is 2057 ± 152 MPa, which corre-
or isostatic pressing of the molybdenum powder into rods and
sponds to results of hardness measurement of P2C sintered molyb-
plates of various geometries and dimensions. The sintering process
denum powders in [12]. Furthermore, the indentation modulus
is carried out in furnaces at high temperatures (typically in the
depends on the porosity. At relative densities up to 95%, an average
range of 1800–2200 °C) and in hydrogen atmosphere over long
value of 297 ± 12 GPa can be calculated which shows an excellent
times (2–3 h) to get densities about 90% of the theoretical density
agreement with Youngs modulus of 305 ± 2 GPa, measured with an
[1,18–24]. For higher densities, hot rolling, extrusion or forging at
ultrasonic scanner at different sintered molybdenum samples, and
temperatures in the range 1200–1500 °C are subsequently used
with values from literature for example in [1,24]. The sintering
[1]. Fig. 3 contains relative densities published in [19], which were
temperature of 2000 °C leads to a softening of the material caused
estimated after 1 h isothermal sintering. For temperatures higher
by the large grain size, but keeping the indentation modulus
than 1400 °C, the comparison shows that the relative densities of
constant.
SPS are comparable or higher than relative densities generated
by isothermal sintering, although the sintering time is very short
at SPS. At lower temperatures isothermal sintered molybdenum 3.3. Microstructure
bodies becomes denser caused by the higher green density due
to higher compaction pressure. Fig. 5 gives an impression of microstructure development
depending on raising sintering temperature at constant external
3.2. Nanohardness pressure of 67 MPa. The cross sections exhibit relatively large
pores and the observed porosity does not agree with that measured
In spite of marginal loading during nanoindentation, porosity by Archimedes technique. The difference is a consequence of prep-
has also strong influence on the values of indentation hardness aration by electrolytic polishing. Caused by the low hardness of
R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557 553

Fig. 5. Scanning electron micrographs of the microstructure of molybdenum at different sintering temperatures, SPS at 67 MPa pressure.

molybdenum (see above), a deformation layer is generated by structure is observed. The sample still exhibits open porosity, the
mechanical grinding and polishing on the cross sections surface. pores are irregularly shaped and exist mainly between the grains.
This deformation layer covers the real microstructure. Addition- The average grain size is larger than the initial average grain size.
ally, pores can be filled during grinding and mechanical polishing. Further increasing of temperature initiates a strong grain growth,
To remove the deformation layer, electrolytic polishing is nec- which was not expected in such a manner at short-time sintering.
essary. The polishing conditions have to be chosen in such a way For sintering temperatures higher than 1600 °C, the porosity is
that the deformation layer is removed completely. However, it is lower than 5%, i.e. closed porosity appears. The pore size decreases
unavoidable then to enlarge existing pores strongly during this and pores exhibit a more and more spherical shape. Pores are dis-
preparation, whereas the etching extends the pores in minor connected from grain boundaries during grain growth and are lo-
way. The single preparation steps are demonstrated in Fig. 6. After cated mainly inside the grains.
mechanical grinding and polishing the surface appears without At the highest temperatures, an agglomeration of pores at grain
contrast, it exhibits small pores and several scratches. At higher boundary triples can be observed. Plane grain faces and grain
magnification the deformation layer is detectable clearly. After boundary angles of about 120° are observed, which is typical for
electrolytic polishing the surface is clean, the scratches are re- polycrystalline metals, since the surface stress of the most grain
moved, grain boundaries become apparent, but the pores are boundaries is similar. In contrast to isothermal sintering of molyb-
strongly enlarged. In contrast, only small material erosion takes denum in [19], no abnormal grain growth was observed.
place during etching. To investigate accurately the relationship between grain growth
The first compact sample was produced with maximum sinter- and temperature, the chord length distribution of the grains was
ing temperature of 850 °C and pressure of 67 MPa. Under these measured in samples sintered with 67 MPa. Fig. 7 shows the rela-
conditions a body of high porosity was obtained, where the shapes tive frequency of chord length depending on sintering tempera-
of origin powder particles are well preserved (Fig. 5). The original ture. A continuous grain growth starts at temperatures higher
shape of particles retains up to sintering temperature of 1100 °C, than 1400 °C. This is closely related to a broadening of the grain
even if the densification is higher. This is in good agreement with spectrum. Also, during isothermal sintering this effect occurs, but
isothermal sintering of molybdenum, where a grain structure at in contrast, it depends on sintering time [10]. When plotting the
1100 °C does not appear [19]. mean chord length vs. the sintering temperature in a log–log scale,
At temperatures of 1400 °C or higher, the original shapes of the two straight lines of different slope can be adapted to the date
powder particles get lost and the formation of a polyhedral micro- points (Fig. 8). The straight lines intersect at the initial temperature
554 R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557

Fig. 6. Preparation steps of cross sections: (a) mechanical grinded and polished; (b) electrolytic polished; (c) etched and (d) deformation layer after mechanical grinding and
polishing. Hardness indentation marks the equal position on sample surface.

Fig. 7. Grain growth of molybdenum by SPS process.

Fig. 9. Grain size, normalized by initial grain size, as a function of the sintering
temperature of our experiments compared with those obtained for isothermal
sintering.

grain size. Up to about 1700 °C the grain sizes of SPS samples are
smaller than those of isothermal sintering. This effect is due to
the shorter sintering times of SPS. Finally, the starting temperature
of grain growth was calculated from the slope of the curves for iso-
thermal sintering (as in Fig. 8). The grain growth starts at temper-
atures between 950 and 1035 °C, which are lower than the SPS
starting temperature of 1330 °C. The huge difference is a conse-
quence of the fast heating up in the SPS process.
The contact of molybdenum powder with the graphite die and
sheets leads to diffusion of carbon inside the samples, although
the spark plasma sintering is a short time process. The micrograph
Fig. 8. Log–log plot of grain growth (67 MPa pressure) vs. sintering temperature. (Fig. 10) shows a sharp transition between carburized region and
uninfluenced molybdenum body. The thickness of the layer de-
pends on temperature and can reach 200 lm at 2000 °C. The layer
for the grain growth. For pure molybdenum the initial temperature grows in a parabolic manner, which allows a rough estimation of
is about 1330 °C, which corresponds to a homologous temperature the activation energy of carbon diffusion in molybdenum. We give
of 0.56 Tm. an estimate of 180 kJ mol 1, which is in agreement with the values
Fig. 9 shows the grain size as a function of the sintering temper- of literature, e.g. in [25]. The carburization kinetics is comparable
ature of our experiments compared with those obtained for iso- with the boriding kinetics in molybdenum under SPS conditions,
thermal sintering [19,20]. The data are normalized by the initial investigated in [26].
R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557 555

Fig. 11. Experimental and corrected piston travel and the corresponding baseline
during sintering of molybdenum powder under a pressure of 67 MPa. The piston
travel, caused by the applied pressure before sintering, was subtracted.

Fig. 10. Carburized region after sintering at 1800 °C.

In the carburized layer, the microhardness rises up to 1300 HV


0.1 and cracks are created. The carbon concentration amounts ca.
7 wt% which is consistent with the carbon content of Mo2C-car-
bide. The results of X-ray phase analyses detected, that the carbu-
rized region was completely converted in the hexagonal Mo2C-
carbide.
The measured diffraction pattern shows a preferred orientation
of the (0 0 2) reflection. Texture measurements verify occurrence
of a (0 0 1) fibre texture modified by a coarse grain structure. The
fibre texture is caused by columnar grain growth, which was ob-
served by light microscopy.

3.4. Densification behaviour


Fig. 12. Corrected piston travel vs. sintering time for different heating rates
The force adjustment of our SPS equipment is starting with a (pressure of 67 MPa).
load of 5 kN. At this point, the motion of piston is recorded during
whole sintering process. The so called ‘‘relative piston travel” con-
Table 2
tains information about linear intrinsic shrinkage of every sample. Values of sample high before sintering.
For interpretation of the records, the experimental data must be
Batch Final height after Final corrected piston Height before
corrected by subtracting the contribution of piston motion caused
sintering lf (mm) travel Dlf (mm) sintering, l0 (mm)
by compression during applied pressure before sintering starts (see
3.1 7.74 0.37 8.11
Fig. 2). This is necessary, because this first compaction depends on
3.2 6.43 0.97 7.40
the applied pressure as well as on the die filling behaviour of pow- 3.3 4.18 2.06 6.24
der varying with every batch. Furthermore, the thermal expansion 3.4 3.89 2.54 6.43
of sample, die and graphite sheets must be subtracted. For this cor- 3.5 4.62 3.79 8.41
rection, which is explained in [6], baselines were determined 3.6 4.65 3.86 8.51
3.7 3.71 3.43 7.14
experimentally. The baselines were recorded with heating rate of
200 K/min, maximal temperature of 2000 °C and pressure of 67
and 57 MPa, respectively. Each baseline can be described by a poly-
nomial, and this is valid for all experimental data recorded at the slopes in parts II and III indicate that the consolidation takes place
same pressure and with the same die. Fig. 11 shows the experi- mainly during fast heating up of SPS.
mental and corrected piston travel and the corresponding baseline Linear shrinkage Dl/l0 can be calculated basing on the corrected
during sintering of molybdenum powder under a pressure of piston travel, where Dl is corrected piston travel depending on sin-
67 MPa. tering time and temperature, respectively. The sample height l0 be-
Fig. 12 gives examples of corrected piston travels depending on fore sintering starts (in our investigation after the cold pressing
different heating rates vs. sintering time. The curves of all samples step) can not be estimated exactly. It can be approximated by
are very similar in their shape but they differ in their slopes the sum of the final sample height lf (after sintering) and the final
depending on the heating rates. On the curves one can observe typ- corrected piston travel Dlf (values see Table 2).
ical points (in Fig. 12 marked by arrows) where the slope changes If this linear shrinkage plotted vs. temperature during heating
considerably. These points are visible in particular for higher sin- up, the curves of all samples, compacted with the same pressure,
tering temperatures (cf. Fig. 12, 1800 °C). By these points, each cover each other with slight deviations (Fig. 13). As pointed out
curve can be divided into parts in the following denoted with Ro- in Section 3.1, the densification of molybdenum by SPS process de-
man numerals. Parts I–III are related to the heating up, the flat part pends strongly on the acting temperature, generated by resistance
IV is controlled by holding time at sintering temperature. The large heating in the powder.
556 R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557

nal pressure and temperature. In comparison with values from the


literature, the densities obtained by SPS exceed those by isother-
mal sintering.
The grain growth during spark plasma sintering was mainly
controlled by the sintering temperature. The grain growth starts
at 1350 °C, which corresponds to a homologous temperature of
0.56 Tm. A strong grain growth occurs at temperatures higher than
1600 °C. Due to the shorter sintering times of SPS the grain sizes of
our samples are smaller than those of isothermal sintering.
Nanohardness and indentation modulus was estimated by
nanohardness testing. It was shown that the porosity strongly
influences both values in spite of the marginal loads used at this
test. An average indentation hardness of 2057 ± 152 MPa and an
average indentation modulus of 297 ± 12 GPa were estimated,
Fig. 13. Corrected piston travel of different sintered batches (pressure of 67 MPa)
which are in agreement with values for bulk molybdenum.
vs. temperature during heating up. Based on the recorded relative piston travel, the linear shrink-
age of molybdenum powder during SPS process can be calculated.
The shrinkage of all batches, compacted with the same external
The parts I up to III marked in Fig. 12 can be observed again and pressure, follow the same curve independent of heating rate. The
their transitions can now be linked to characteristic values of tem- main part of compaction takes place during fast heating up by SPS.
perature, shrinkage, porosity and mean grain size, as demonstrated In summary, our investigation shows that the best results of
in Fig. 13. The reason for the small shift of the curves of Fig. 13 is molybdenum powder spark plasma sintering are achieved at
not clear. 1600 °C and 67 MPa. The optimal sintering temperature corre-
sponds to that of Plasma Pressure Compaction, but a higher exter-
nal pressure is necessary for SPS.
4. Discussion

Acknowledgments
As the densification process is thermally activated, a sigmoid
progression of sintering curve was expected, comparable with such
The authors are grateful for the financial support of this work by
of isothermal sintering [27,28], where in our curves the part I
Dr. Erich Krüger Stiftung and for the abandonment of the molybde-
exhibits a linear relationship between Dl/l0 and temperature and
num powder by Dr. A. Hoffmann, Plansee Metall GmbH Reutte.
a sharp transition to second part starting at 900 or 950 °C, respec-
tively. This transition temperature is consistent with 0.4 Tm of Mo,
References
the temperature, where appreciable volume diffusion starts in so-
lid materials. The end of this part is related to a small shrinkage of [1] Plansee high performance materials, Molybdenum – material properties and
5% and a high porosity of 40%. The densification rate is low. As applications. The Official Website of the Plansee Group, <http://
shown in Section 3.3, grain growth and change in particle shape www.plansee.com/lib/Molybdenum.pdf> [accessed 15.3.2010].
[2] Tokita M. Mechanism of spark plasma sintering and its application to ceramics.
were not observed. Nyu Seramikkusu 1997;10:43–54. Alternative: SPS SYNTEX INC, What’s SPS?.
Part II is characterised by the highest densification rate of sin- The Official Website of the SPS SYNTEX INC, <http://www.scm-sps.com/
tering process. The shrinkage achieves more than 30% and the index.html> [accessed 15.3.2010].
[3] Omuri M. Sintering, consolidation, reaction and crystal growth by the spark
porosity decreases to 15%, where the micrographs do not show
plasma system (SPS). Mater Sci Eng A 2000;287:183–8.
any change in particle shape up to 1100 °C. Nevertheless, the den- [4] Chen W, Anselmi-Tamburini U, Garay JE, Groza JR, Munir ZA. Fundamental
sity increases and the number and sizes of contact areas between investigations on the spark plasma sintering/synthesis process. I. Effect of dc
pulsing on reactivity. Mater Sci Eng A 2005;394:132–8.
particles are enlarged. Furthermore, the particle strength decreases
[5] Anselmi-Tamburini U, Gennari S, Garay JE, Munir ZA. Fundamental
further, which makes particle rotation and sliding easier. A phe- investigations on the spark plasma sintering/synthesis process. II. Modelling
nomenological description of the consolidation process during of current and temperature distributions. Mater Sci Eng A 2005;394:139–48.
parts I and II is an open problem. [6] Anselmi-Tamburini U, Garay JE, Munir ZA. Fundamental investigations on the
spark plasma sintering/synthesis process. III. Current effect on reactivity.
The grain growth starts with the beginning of the third part at Mater Sci Eng A 2005;407:24–30.
1330 °C. At first the porosity amounts 10%, which marks the tran- [7] Chaim R. Densification mechanism in spark plasma sintering of
sition to closed porosity. During the third part the porosity reduces nanocrystalline ceramics. Mater Sci Eng A 2007;443:25–32.
[8] Munir ZA, Anselmi-Tamburini U, Ohyanagi M. The effect of electric field and
to 5% and the densification rate decreases continuously. pressure on the synthesis and consolidation of materials: a review of the spark
The micrographs (Fig. 5, 1600 °C) show that the pores are plasma sintering method. J Mater Sci 2006;41:763–77.
rounded and the grain boundaries are separated from pores. Both [9] Orrú R, Licheri R, Locci AM, Cincotti A, Cao G. Consolidation/synthesis of
materials by electric current activated/assisted sintering. Mater Sci Eng R
processes are driven mainly by diffusion. 2009;63:127–287.
In part IV, where sintering temperature is constant, the densifi- [10] Zhang ZH, Wang FC, Wang L, Li SK. Ultrafine-grained copper prepared by spark
cation is comparable to isothermal sintering process. The material plasma sintering process. Mater Sci Eng A 2008;476:201–5.
[11] Srivatsan TS, Ravi BG, Nauka AS, Riester L, Petraroli M, Sudarshan TS. The
transport is supported by the applied pressure and by the electric
microstructure and hardness of molybdenum powders consolidated by plasma
field, which was demonstrated in [6,8]. pressure compaction. Powder Technol 2001;114:136–44.
[12] Srivatsan TS, Ravi BG, Petraroli M, Sudarshan TS. The microhardness and
microstructural characteristics of bulk molybdenum samples obtained by
5. Conclusions consolidating nanopowders by plasma pressure compaction. Int J Refract Met
Hard Mater 2002;20:181–6.
[13] Yamauchi A, Yoshimi K, Kurokawa K, Hanada S. Synthesis of Mo–Si–B in situ
The results of this investigation demonstrate that molybdenum composites by mechanical alloying. J Alloys Compd 2007;434–435:420–3.
powder can be successfully consolidated by spark plasma sintering [14] Impermeable sintered metal materials and hardmetals – determination of
in very short times. A relative density of 95% is achieved at sinter- density [ISO 3369: 1975]. German version DIN ISO 3369; 1990.
[15] Sintered metal materials, excluding hardmetals – permeable sintered metal
ing temperatures of 1600 °C and an external pressure of 67 MPa. It materials – determination of density, oil content and open porosity [ISO 2738:
can be concluded that the density is primarily controlled by exter- 1999]. German version EN ISO 2738; 2000.
R. Ohser-Wiedemann et al. / Int. Journal of Refractory Metals & Hard Materials 28 (2010) 550–557 557

[16] Metallic materials – instrumented indentation test for hardness and materials [23] Fan J, Lu M, Cheng H, Tian J, Huang B. Effect of alloying elements Ti, Zr on the
parameters – part 1: test method [ISO 14577-1: 2002]. German version EN ISO property and microstructure of molybdenum. Int J Refract Met Hard Mater
14577-1; 2002. 2009;27:78–82.
[17] Database PDF-2 of the International Centre for Diffraction Data; 2006. [24] Brooks K. New ways to make moly as it enters nano-phase production. Met
[18] Garg P, Park SJ, German RM. Effect of die compaction pressure on densification Powder Rep 2004;59:18–21.
behaviour of molybdenum powders. Int J Refract Met Hard Mater 2007;25: [25] Landolt-Börnstein – diffusion in solid metals and alloys. Group III. Condens
16–24. Matter, vol. 26. Berlin Heidelberg New York: Springer; 2006.
[19] Majumdar S, Raveedra S, Samajdar I, Bhargava P, Sharma IG. Densification and [26] Yu LG, Khor KA, Sundararajan G. Boride layer growth kinetics during boriding
grain growth during isothermal sintering of Mo and mechanically alloyed Mo of molybdenum by the spark plasma sintering (SPS) technology. Surf Coat
TZM. Acta Mater 2009;57:4158–68. Technol 2006;201:2849–53.
[20] Kim G, Kim HG, Kim D, Oh S, Suk M, Kim YD. Densification behaviour of Mo [27] German RM. Powder metallurgy and particulate materials
nanopowders prepared by mechanochemical processing. J Alloys Compd processing. Princeton, NJ: Metal Powder Industries Federation; 2005.
2009;469:401–5. [28] Salmang H, Scholze H. Keramik. Berlin Heidelberg New York: Springer; 2007.
[21] Schatt W, Wieters KP, Kieback B. Pulvermetallurgie. 2nd ed. Berlin Heidelberg,
New York: Springer; 2007.
[22] Riedel H, Svoboda J, Huber K, Plankensteiner A. On physico-chemical
mechanisms in desoxidation and sintering of molybdenum. In: Proceedings
of the 17th international plansee seminar 2009;3:WS 5/1–WS 5/12.

You might also like