You are on page 1of 7

Materials Science and Engineering A 521522 (2009) 322328

Contents lists available at ScienceDirect

Materials Science and Engineering A


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

Understanding the mechanical properties of hardmetals through mechanical


spectroscopy
D. Mari
Ecole Polytechnique Fdrale Lausanne, IPMC, Station 3, 1015 Lausanne, Switzerland

a r t i c l e

i n f o

Article history:
Received 3 June 2008
Accepted 2 October 2008
Keywords:
Tungsten carbide
Cobalt
Nickel
Internal friction
Cermet
Cemented carbide

a b s t r a c t
Composites made of transition metal carbides, with a ferrous metal binder, commonly dened as hardmetals, are used for cutting tool applications. We investigate the mechanical behavior of the two main families
of hardmetals: the cemented carbides, constituted of WCCo, and the cermets made of a combination of
TiCTiN(Co,Ni). The comparative analysis of internal friction spectra of both hardmetal families as well
as the analysis of spectra of the ceramic phases alone (after binder etching) allows associating different
relaxation peaks with specic phases or grain boundaries. The mechanical properties of hardmetals are
controlled by the binder. Three temperature domains can be dened. Increasing the temperature binder
microplasticity occurs rst, then binder segregation at the ceramic grain boundaries prepares the way for
inltration and extensive deformation by grain boundary sliding.
2009 Elsevier B.V. All rights reserved.

1. Introduction
Composites made of transition metal carbides with a ferrous
metal binder are widely used for the manufacturing of cutting
tools. They are commonly dened as hardmetals. Among them, two
main families of materials can be distinguished: the cemented carbides, mainly constituted of WCCo, and the cermets made of a
combination of TiCTiN(Co,Ni) with other carbides [1]. The main
qualities required for a cutting tool are a good toughness and a
good resistance to deformation at high temperature. The cemented
carbides have a very high toughness and good hardness. However,
the performances of such materials are rapidly deteriorated when
high cutting speed and high temperatures are involved because
WCCo has a poor chemical resistance and undergoes plastic deformation at high temperature [2]. The cermets have in general a
better resistance to high temperature deformation but toughness
is lower than in cemented carbides. The morphology of hardmetals
is well described by a structure made of two interpenetrated skeletons: the ceramic and the metal ones. The ceramic skeleton is an
arrangement of joined single-crystal carbide or carbonitride grains.
In WCCo, grains are faceted with triangular or rectangular shape
revealing the hexagonal structure of WC. In TiCN-based composites,
the grains have a core-rim structure where the core corresponds
to pure TiCN and the rim to (Ti,M)CN precipitated during solidliquid phase sintering [35]. M represents a transition metal carbide

Tel.: +41 21 693 4473; fax: +41 21 693 4470.


E-mail address: daniele.mari@ep.ch.
0921-5093/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2008.10.064

former. Mechanical spectroscopy is a powerful tool to analyse the


properties and the evolution of these composite materials since its
spectroscopic character allows the separation of the microplasticity
mechanisms that occur in each phase. A wide range of hardmetals
has been studied to date [6,7]. The analysis of internal friction (IF)
spectra is complex because of the wide range of phenomena that
may occur in such composite materials. The purpose of this paper
is to show a unied view of the mechanism underlying the internal
friction spectra. Only the comparison of the spectra obtained in different materials allows the attribution of each internal friction peak
to a specic phase and mechanism in the composite. Such analysis is extremely fruitful if compared with macroscopic mechanical
properties and leads to the understanding of the evolution of the
hardmetals with temperature.
2. Experimental techniques and materials
Different hardmetals are presented in this study (Table 1): two
grades of cemented carbides containing 10 and 18 vol%Co and three
cermets with nickel and cobalt binder. Samples were produced by
sintering at a typical temperature of 1673 K for WCCo and 1723 K
for the cermets. Fig. 1 shows the typical morphology of a cemented
carbide and of a cermet. The samples for internal friction measurements were spark-cut at the nal dimension of 1 4 4060 mm3 .
In order to separate the effect of the ceramic skeleton from that of
the binder, the metal phase was extracted from the IF samples by
chemical etching with boiling concentrated HCl. Mechanical spectroscopy was performed in a sub-resonant, forced, inverted torsion
pendulum [6]. Measurements were carried out at constant heating

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

323

Table 1
Composition of the alloys presented in this work. Phase fractions are given in at%. The
grade denominations instead refer to vol%. In TiCN-based cermets, if not specied
the ratio C/N is 30/70.
Alloy

Co/Ni

Ti

Mo

WC18Co
WC10Co
TiCN6Mo18Co
TiCN13Mo7Ni
TiCN5WC10Co
(TiC0.5 N0.5 )6Mo6Co
TiCN6Mo6Co

16.93
9.71
9.13
6.15
9.13
10.47
5.80

41.53
45.15
0.00
0.00
3.92
0.00
0.00

0.00
0.00
38.96
41.37
42.08
45.39
45.21

0.00
0.00
3.65
7.40
0.00
1.72
1.70

0.00
0.00
11.69
12.41
16.41
22.82
13.76

41.53
45.15
26.99
32.66
28.46
24.22
33.53

rate (1 K/min) and at a frequency of 1 Hz in a temperature range


from 300 K to 1500 K, and isothermally in a frequency range from
1 mHz to 10 Hz. In forced mode, the IF is measured by the tangent
of phase lag: tan(). The spectra obtained both from frequency and
temperature scans are deconvoluted into several, possibly broadened, Debye peaks, each of which is described by a function:
=

()
1 + (2  2 )

0<1

(1)

where  is the relaxation strength, the angular frequency,  the


relaxation time and 1/ the broadening factor [8]. For a thermally
activated mechanism, the relaxation time depends on temperature
according to an Arrhenius law:
 = 0 eEact /kT ,

(2)

where Eact is the activation energy, k is Boltzmanns constant and  0


is a constant corresponding to the inverse of the attempt frequency
0 .
3. Results

Fig. 2. Internal friction spectra in a WC18 vol%Co measured at 1 Hz as a function


of temperature upon heating. Curve A represents the as received state. Curve B is
observed after heating above 1200 K and is stable. Curve C is obtained from status B
after annealing for 72 h at 1000 K. Curve D represents the internal friction of the WC
skeleton.

is stable during the following heating-cooling cycles. However, if


the sample undergoes a long annealing at 1000 K, a new spectrum
characterized by a peak at 900 K is obtained. Curve C clearly shows
a transition at 1050 K between the two spectra. If the sample is
heated above 1050 K the curve B is obtained again. The spectrum of
the WC skeleton does not show any of the above features but some
increasing background above 1100 K.
A typical spectrum of TiCN6Mo18Co is shown in Fig. 3. It shows
two peaks respectively at 1000 K and 1200 K and a background. The
rst peak PT0 is also found in the skeleton spectrum while other
features are less clearly visible. On the second heating, the peak
PT0 has disappeared making PT1 more visible.

3.1. Base spectra of WCCo and TiCNCo


3.2. Deconvolution of the low temperature part of the IF spectra
A characteristic spectrum of WC18 vol%Co is presented in Fig. 2.
In the rst heating (curve A), only a bump at 1150 K can be clearly
distinguished. A peak PW1 is better dened in the second heating with maximum at 1100 K (curve B). The spectrum of curve B

In order to better extract the information from the IF spectra,


the high temperature part of them can be assimilated to an exponential background and subtracted. Fig. 4 shows the spectrum of

Fig. 1. Grain structure of (a) WC18 vol%Co cemented carbide and (b) TiCN-5 vol%WC10 vol%Co cermet. The cobalt is bright in the cermet and darker in the WCCo. In the
cermet, the typical core-rim structure can be observed. The white regions near the core represent the inner rim formed during solid state sintering.

324

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

Fig. 3. Internal friction spectrum of TiCN6Mo18Co measured at 1 Hz upon heating. In


the rst heating, two peaks PT0 and PT2 can be distinguished. In the second heating,
PT0 is missing and reveals the presence of PT1 . PT0 is also clearly present in the
TiMoCN skeleton obtained by etching the binder.

WC18Co (curve C) after subtraction of the background. We can


observe more clearly the transition from the low temperature state
to the high temperature one at 1050 K. In order to drive the analysis
of the spectrum, a peak is tted to the low temperature part of the
spectrum and two are tted to the high temperature state. The peak
PC has been introduced in the deconvolution based on the fact that
at that temperature the magnetic transition of the cobalt in WCCo
is expected [9]. However, such a transition does not correspond with
the peak shape of a Debye relaxation.
Fig. 5 shows the deconvolution of the spectrum of TiCN6Mo18Co
after background subtraction. Four peaks numerically ordered as a
function of their temperature maxima are tted to the experimental
curve. The spectrum of the skeleton is equally reported. The t of
this curve shows that the peaks PT0 , PT2 and PT3 are also present
in the skeleton (even if with lower amplitude) while PT1 is absent.
Some of the features observed in TiCNMoCo are equally found
in cermets with a Ni binder as shown in Fig. 6. The peaks PT2 and PT3
are also found in TiCN13Mo7Ni, which conrms that they belong
to the carbide skeleton.
All peaks PT0 , PT2 and PT3 did show an amplitude that increases
with the binder content [10], which therefore may partially
account for their decrease when the measurement is performed
on the skeleton only. In Fig. 7, the spectra of a high nitrogen

Fig. 4. Deconvolution of the internal friction spectrum (curve C in Fig. 2) after high
temperature background subtraction. A transition between two peaks PW1 at low
temperature and PW1 at high temperature can be observed. Another peak PC is tted
in correspondence with the magnetic transition.

Fig. 5. Deconvolution of the internal friction spectrum of TiCN6Mo18Co and of its


respective TiMoCN skeleton after high temperature background subtraction during
rst heating. Four peaks can be tted in the complete material but PT1 is absent in
the skeleton.

Fig. 6. Deconvolution of the internal friction spectrum of TiCN13Mo7Ni and of its


respective TiMoCN skeleton after high temperature background. Three peaks are
tted to the complete material curve. They are evidently also present in the skeleton.

Fig. 7. Deconvolution of the spectra of (curve 1) (TiC0.5 N0.5 )6Mo6Co and (curve 2)
(TiC0.3 N0.7 )6Mo6Co after background subtraction. The peak PT0 is much higher in
the high nitrogen alloy. After rst heating (curve 3) PT0 has disappeared; PT1 and
PT2 can be distinguished.

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

325

(TiC0.5 N0.5 )6Mo6Co and high carbon (TiC0.7 N0.3 )6Mo6Co cermet
are presented. Their comparison shows a strong dependence of PT0
and PT2 on nitrogen content in TiCN. Upon the rst heating, PT0 is
clearly higher in (TiC0.5 N0.5 )6Mo6Co. In the second heating, where
PT0 has disappeared, we can notice that also PT2 is much higher in
the high nitrogen sample.
3.3. Deconvolution of the high temperature part of the IF spectra:
frequency scans
The high temperature part of the spectra is not accessible by
temperature scans at frequencies in the range of 1 Hz typical of
free pendula. Frequency scans performed with forced pendula add
information about high temperature mechanisms. Fig. 8a shows the
deconvolution of the spectrum of WC10Co in isothermal frequency
scans. A low frequency the background is subtracted. Two relaxation peaks PW2 and PW3 are well dened. Their tted shape is
shown for one temperature. The amplitude of PW2 and PW3 varies
in parallel as a function of the temperature attaining a maximum at
about 1350 K (Fig. 8b). The same peaks also were found in the WC
skeleton at much lower amplitude. In reality, the low frequency

Fig. 9. (a) High temperature frequency scan in TiCN5WC10Co. At very low frequency,
a high temperature peak is outlined PTht. (b) After subtraction of the tted peak,
another peak PT4 is clearly appearing between 0.1 Hz and 1 Hz. The amplitude of
this peak is increasing as a function of temperature in a reversible way. For clarity,
only two curves measured on cooling at 1425 K and 1475 K are reported.

Fig. 8. Deconvolution of the frequency scan spectrum of WC10Co. (a) Two peaks
PW2 PW3 can be distinguished. They are thermally activated and (b) their amplitude
varies as a function of temperature.

part of the spectrum is a true relaxation peak as demonstrated in a


detailed study [11].
In Fig. 9a and b frequency scans performed on a cermet containing W instead of Mo and with 10 vol%Co binder are presented.
The morphology is that of classical cermets with W forming the
rim of (TiW,CN) cubic phase. The raw data are presented in Fig. 9a.
By looking at the curves collected at the highest temperatures, the
right ank of a peak can be perceived. Therefore it can be argued
that the high temperature background in these materials may be
in reality a relaxation peak (PTht). The comparison with WCCo,
where a relaxation peak is found, supports this hypothesis. After
the t of the hypothetical high temperature peak in the cermet,
another peak can be observed in frequency scans (Fig. 9b). This peak
PT4 is characterized by an amplitude increasing with increasing the
temperature. This amplitude dependence is perfectly reversible as

326

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

Table 2
Fit parameters of the peaks found in TiCN-based cermets. Activation energies used in the t are taken from isothermal measurements.
/2

Temperature (1 Hz)

Activation energy (eV)

Attempt frequency, 0 (Hz)

Broadening 1/

PT0
PT0 in skeleton

28 10
2 103

920
920

2.2
2.2

10
1013

1.3
1.9

PT1

24 103

980

1.6

109

1.8

13

PT2
PT2 in skeleton

48 10
1 103

1150
1150

2.1
1.7

10
109

1.6
1.6

PT3
PT3 in skeleton

1 103
0.5 103

1500
1500

2.7
2.3

1012
1010

1.3
1.2

PT4
PT4 in skeleton

060 103
3 103

1500
1400

4.0
3.4

1015
1013

1.3
1.0

PTht

0.8

>1700

5.3

1017

3.0

10

Table 3
Fit parameters of the peaks found in WC-based cemented carbides. Activation energies used in the t are taken from isothermal measurements.
/2

Temperature (1 Hz)
3

Activation energy (eV)

Attempt frequency, 0 (Hz)

Broadening 1/

PW1
PW1

15 10
13 103

1080
1000

2.5
2.5

10
1012

1.3
2.5

PW2
PW2 in skeleton

120 103
4 103

1330
1330

2.0
2.6

1010
1010

1.4
1.5

PW3
PW3 in skeleton

115 103
5 103

1440
1440

4.5
4.9

1016
1018

1.2
1.2

shown in some of the curves reported on cooling. On the other hand,


the peak PT4 is also found in TiWCN skeletons but at much lower
and constant amplitude tg() = 3 103 .
The thermal activation parameters of TiCN-based cermets are
given in Table 2 and those related to WC-based cemented carbides
are reported in Table 3.
4. Discussion
Internal friction can contribute quite interestingly to a global
view of the evolution of the mechanical properties of hardmetals with temperature. Cemented carbides (WC-based) and cermets
(TiCTiN-based) shows striking similarities in the internal friction
spectra as a function of temperature. We can analyse the change of
the mechanical behavior in different temperature domains characterized by relaxations specic to each phase or grain boundaries.
4.1. Binder microplasticity: peaks PW1 and PT1
At about 1100 K, all hardmetals containing cobalt show a characteristic peak. This peak is stable in cermets where the Co binder
contains Mo as a solute (Fig. 5) while it shows two different states
when the solute is tungsten. The peak is absent in skeleton samples
and should be unambiguously attributed to the cobalt binder. In the
present study, a careful analysis of the spectrum recorded during
the transition shows indeed that the low temperature peak PW1
is replaced by another at higher temperature. They should therefore be attributed to the same defect. This transition in the binder
behavior was rst evidenced by Schaller et al. [12] and attributed
to a change in solute W concentration together with a change in
the dislocation structure from partial dislocations to recombined
perfect dislocations [6,13]. Effectively, it is well documented that
the cobalt in WCCo has an fcc () structure with a stacking fault
network that highly increases its yield stress [6]. The stabilization
of the low temperature peak requiring a long annealing at 1000 K
suggests a precipitation mechanism. The precipitation of WCo3 has
been evoked in [12] but the WCo phase diagram [14] conrmed
by unpublished results by the author shows that the dissolution
of this compound could be expected only above 1170 K. This is not

12

consistent with the present transition temperature of 1050 K. In this


paper, we propose a new interpretation of the peaks appearing in
the cobalt phase of the WCCo based on recent experimental proofs
of the existence of a spinodal decomposition in the CoW phase
diagram. In WCCo, upon the rst heating, a sharp peak appears
clearly at 1170 K (Fig. 2). It should be attributed to the magnetic
transition at the Curie temperature of the cobalt as shown by Gardon [9]. However, magnetic effects can produce a decomposition of
CoW that was rst proposed in the calculation of phase diagram by
Fernndez Guillermet [15]. A ferromagnetic W-poor and a paramagnetic W-rich phase should form in fcc cobalt together with WCo3 .
Recently, stberg et al. [16] have given experimental evidence of
the presence of bands in the cobalt phase of a cermet with periodically uctuating W concentration. Such concentrations were those
expected from the miscibility gap at about 1000 K. We suggest that
the annealing at 1000 K promotes the spinodal decomposition. The
diffusion of W is however very sluggish and long annealing times
are needed to obtain such microstructure. The transition temperature of 1050 K, observed in Figs. 2 and 3 would then correspond to
the solubility limit of the W-rich phase. The initial W solute content
in the cobalt can be assessed by the dependence of the Curie temperature on the W concentration in homogeneous solution [15]. The
temperature of the magnetic transition peak in curve A indicates a
W content of about 6 at% that corresponds rather well with that
found in stbergs work. We expect therefore a similar microstructure. The internal stress effect are however neglected. PW1 could
be attributed to dragging of W atoms by dislocations. The activation energy of 2.5 eV corresponds very well with diffusion energy
of W into cobalt [17]. In the low temperature state, where the spinodal decomposition is present, the W-rich zones behave like strong
pinners for dislocations that are free to move in the W depleted
zones. According to [16], these zones may have a size of about 10 nm.
When the W-rich phase is dissolved, the dislocations will have a
free length similar to the cobalt mean free path i.e. 100 nm. Taking
the classical model for dragging [18], the relaxation time should be
proportional to L2 /d, where L is the dislocation free length and d is
the distance between dragged point defects. If there we consider
a change of L by a factor 10 and an activation energy of 2.5 eV, the
peak would shift from 950 K to 1120 K in agreement with Fig. 4. The

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

large broadening of the peak PW1 can be justied by the fact that
in reality the spinodal decomposition has a quite irregular distribution. Moreover, there may be some regions where it is even not
present yielding dislocations with a larger free length. The hypothesis formulated in this work also justies the observation of a critical
behavior activation energy of the peak PW1 with a critical temperature estimated at 1073 K [19]. The model described in Ref. [19] is
essentially correct even if the phase transformation that should be
considered is not the hcp to fcc transformation but the solution
of the spinodal. The high temperature stable peak PW1 should be
a dragging peak than PW1 as conrmed by the activation energy
and attempt frequency. An important consequence of this analysis
is that the spinodal decomposition of the WCo alloy can give a substantial contribution to cobalt hardening in the cemented carbides
and therefore increase their mechanical resistance to deformation.
In the TiCNMoCo alloys, the peak PT1 has never shown a
behavior similar to PW1 even if from the phase diagram an equivalent spinodal decomposition could be expected [20]. This peak
should be attributed to Mo dragging by dislocations [10].
4.2. Towards grain boundary sliding: point defect relaxations at
grain boundaries
At intermediate temperatures, a number of relaxation peaks
appears both in cermets and in cemented carbides. At about 980 K,
the peak PT0 is observed in Co-based cermets. It characterizes the
as sintered structure and disappears after the rst heating above
1000 K. This peak is generally sensitive to material chemistry. It
increases with the cobalt content and, as shown in Fig. 7, it depends
strongly on the nitrogen content in the ceramic phase. It was shown
previously [10] that when the peak is present a thin cobalt lm
can be detected at grain boundaries. Therefore, the presence of PT0
witnesses the ability of cobalt to inltrate the grain boundaries
in a more or less massive way. On the other hand, Fig. 7 shows
that the peak PT2 is highly inuenced by nitrogen content and
in another paper [10] the dependence of PT2 PT3 on cobalt content was demonstrated. The similarity of the behavior of PT0 with
the PT2 PT3 pair suggests a different interpretation of these two
peaks with respect to former literature. The presence of PT2 PT3
has been associated with the movement of dislocations in the TiCN
grains controlled by a mechanism equivalent to the Bordoni peak
for the ceramic phase [6,21]. This interpretation was justied by
the observation of dislocation multiplication when cermets are
deformed in the temperature range where PT2 PT3 are observed at
1 Hz, and by the dependence of the peak amplitude on the binder
content. This last effect was supposed to be induced, as the Bordoni peak, by internal stresses [22]. In fact, such stresses increase in
the ceramic phase when the binder content increases. However, as
demonstrated by diffraction measurements, the residual stresses
are almost completely relaxed above 1200 K [23] where PT2 PT3
appear. As shown in [10], PT2 PT3 are also quite sensitive to the Mo
content in the ceramic. The sensitivity to the ceramic composition
and to the cobalt content suggests that the process involves both
constituents. The process must however be located in the ceramic
since it is present when the cobalt is removed or with a Ni binder.
The most likely hypothesis is that grain boundaries are involved
and that the process is related to cobalt (or nickel) segregation at
the grain boundaries. As shown in Fig. 7, also nitrogen has an important effect. Nitrogen is known to modify the Mo activity to favor its
dissolution in the binder [24]. Therefore, we propose that both PT0
and the peaks PT2 PT3 involve diffusion processes at the TiCN grain
boundaries enhanced by the solution of Mo or W into the binder.
It is tempting to draw a parallel with peaks PW2 and PW3 in the
cemented carbides. Also these two peaks are always detected in
parallel and have similar activation energies. They appear at higher
temperature though. The peaks PW2 PW3 show a varying ampli-

327

tude with a maximum at 1350 K. A stable single layer of Co has


been found in WCCo [25] at WCWC interfaces. Recent calculations show that such layer stabilizes the grain boundary against
further inltration but the situation may change with temperature
and stress. It interesting to notice that PW2 PW3 tend to disappear when the peak related to grain boundary inltration by cobalt
[11] starts to appear. This would indicate that when complete inltration of the grain boundary starts the peaks that are related to
segregation disappear.
Also the peak PT4 shows a varying amplitude but in this case it
practically increases linearly with temperature starting from 1350 K
(Fig. 9b). Moreover, each isothermal state seems at equilibrium
and reversible. The effect is clearly related with the presence of
cobalt since when cobalt is removed the peak has a stable amplitude (Table 2). The corresponding peak found in the skeleton [23]
is practically a Debye peak showing that the mechanism originating PT4 should be a point defect relaxation. PT4 could be related to
the progressive segregation of cobalt at the grain boundaries. However, in isothermal measurements, PT4 is present in parallel with
the high temperature background (or presumably PTht). Assuming
that PTht is due to high temperature inltration of grain boundaries by the cobalt, it could be argued that either only some grain
boundaries are inltrated or that PT4 pertains only to some grain
boundary congurations.
In essence, the presence of different peaks probably located at
the grain boundaries suggests that their modication with temperature occurs in different stages. Different levels of grain boundary
segregation with phase transitions according to the number of segregated layers have been proposed by Creuze et al. [26,27]. These
authors show that the grain boundary structure evolves with segregated atoms involving progressively 1 to several layers close to the
grain boundary. At a certain level of saturation and temperature, a
wetting transition may occur with phase separation. We conclude
that all peaks found in the hardmetals at intermediate temperatures
reveal different states of intermediate equilibrium of grain boundaries related with binder segregation. These states prepare the high
temperature inltration that leads to the extensive deformation of
these composites [28].
Wetting of the grain boundaries is likely to occur at high temperature. This effect has been clearly evidenced in cermets [29]
and cemented carbides [6,30] and is directly related to extensive
deformation of these composites at high temperature.
5. Conclusions
Hardmetals show a complex internal friction spectrum due to
phase interaction.
The analysis of temperature scans in parallel with isothermal
frequency scans demonstrates in a reliable way the presence of a
multiplicity of relaxation peaks.
In WCCo, a relaxation peak attributed to Co interaction with
solute W has been identied. The abrupt change of its shape gives
further evidence of a spinodal decomposition of the WCo alloy
constituting the binder.
At intermediate temperature, different point defect located at
the WC or TiCN grain boundaries reveal the progressive change of
these grain boundaries due to binder segregation.
At high temperature, the wetting of the binder of the grain
boundaries produces a grain boundary sliding peak both in the
cemented carbides and in the cermets.
References
[1] D. Mari, Encyclopedia of Materials, Elsevier Science Publisher, Amsterdam,
2001, p. 11181123.
[2] D. Mari, D.R. Gonseth, Wear 165 (1993) 917.

328

D. Mari / Materials Science and Engineering A 521522 (2009) 322328

[3] D.I. Chun, D.Y. Kim, J. Am. Ceram. Soc. 76 (1993) 20492052.
[4] H.-O. Andrn, U. Rolander, P. Lindahl, Int. J. Refract. Metals Hard Mater. 12
(19931994) 107113.
[5] J. Zackrisson, U. Rolander, H.-O. Andrn, Metall. Mater. Trans. A 32 (2001) 8594.
[6] D. Mari, S. Bolognini, G. Feusier, T. Viatte, W. Benoit, Int. J. Refract. Metals Hard
Mater 17 (1999) 209225.
[7] D. Mari, S. Bolognini, T. Viatte, W. Benoit, Int. J. Refract. Hard Metals 19 (2001)
257265.
[8] R.M. Fuoss, J.G. Kirkwood, J. Am. Chem. Soc. 63 (1941) 385394.
[9] M. Gardon, Etude par Spectromtrie Mcanique de lEvolution Structurale
Haute Temprature du Composite WC-Co, Ph.D. Thesis INSA, Lyon 1993.
[10] S. Bolognini, G. Feusier, D. Mari, T. Viatte, W. Benoit, Int. J. Refract. Metals Hard
Mater. 16 (1998) 257268.
[11] D. Mari, K. Buss, W. Benoit, Mater. Sci. Eng. A 442 (2006) 179183.
[12] R. Schaller, J.J. Ammann, C. Bonjour, Mater. Sci. Eng. A 105106 (1988) 313321.
[13] R. Schaller, D. Mari, M. Maamouri, J.J. Ammann, J. Hard Mater. 3 (1992) 351362.
[14] H. Okamoto, J. Phase Equilib. Diffus. 29 (2008) 119.
[15] A. Fernndez Guillermet, Metall. Trans. A 20A (1989) 935956.
[16] G. stberg, B. Jansson, H.-O. Andrn, Scripta Mater. 54 (2006) 595598.
[17] A. Davin, V. Leroy, D. Coutsouradis, L. Habraken, Memoires Scientiques de la
Revue de Mtallurgie 60 (1963) 275283.

[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]

K. Lcke, A.V. Granato, J. Phys. 42-C5 (1981) 327337.


J.-J. Ammann, R. Schaller, J. Alloys Compd. 211212 (1994) 397401.
A. Davydov, U. Kattner, J. Phase Equilib. Diffus. 24 (2003) 209211.
S. Bolognini, D. Mari, T. Viatte, W. Benoit, Advances in Mechanical Behaviour,
Plasticity and Damage, Elsevier Science, Oxford, 2000, p. 777782.
G. Fantozzi, C. Esnouf, W. Benoit, I. Ritchie, Progr. Mater. Sci. 27 (1982)
311451.
K. Buss, High temperature deformation mechanisms of cemented carbides and
cermets, Ph.D. Thesis EPFL, Lausanne, 2004.
P. Lindahl, U. Rolander, H.-O. Andrn, J. Hard Mater. 3 (1992) 259267.
A. Henjered, M. Hellsing, H.-O. Andrn, H. Nordn, Mater. Sci. Tech. 2 (1986)
847855.
F. Berthier, J. Creuze, R. Ttot, B. Legrand, Phys. Rev. B 65 (2002), 195413-195411195417.
J. Creuze, F. Berthier, R. Ttot, B. Legrand, Surf. Sci. 553 (2004) 168180.
G. stberg, K. Buss, M. Christensen, S. Norgren, H.-O. Andrn, D. Mari, G. Wahnstrm, I. Reineck, Int. J. Refract. Metals Hard Mater. 24 (2006) 135144.
S. Bolognini, G. Feusier, D. Mari, T. Viatte, W. Benoit, Int. J. Refract. Metals Hard
Mater. 21 (2003) 1929.
G. stberg, H.-O. Andrn, Metall. Mater. Trans. A 37 (2006) 14951506.

You might also like