Investigation of The Sliding Contact Properties of WC-Co Hard Metals Using Nanoscratch Testing

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Wear 263 (2007) 1602–1609

Investigation of the sliding contact properties of


WC-Co hard metals using nanoscratch testing
Siphilisiwe Ndlovu ∗ , Karsten Durst, Mathias Göken
Materials Science and Engineering, University of Erlangen-Nürnberg, D-91058 Erlangen, Germany
Received 15 August 2006; received in revised form 10 November 2006; accepted 16 November 2006
Available online 23 May 2007

Abstract
Nanoscratch tests were performed on a range of WC-Co hard metals with varying cobalt content and WC grain size using a Nanoindenter XP.
Single and multiscratch tests were conducted with various loads.
The scratch friction coefficient for all the samples was found to be approximately 0.4 and was observed to fluctuate due to the hard and ductile
phases in the material. The scratch width and depth were found to increase with increasing load for single scratches. Multiscratching with a constant
load resulted in the widening and deepening of the scratches at each load with accumulative damage occurring as more tests were performed.
Damage was mainly attributed to a brittle mechanism occurring via the formation and interaction of subsurface cracks and the deformation of WC
grains via slip. A deformed layer was formed on the surface of several of the hard metals during multiscratching and this was found to contain
WC fragments which were formed during testing. Cobalt extrusion also took place and in the case of the 6 and 6.5% cobalt samples, led to the
subsequent loss of loosely anchored WC grains. The scratch depth and width were found to increase linearly with load with more severe grain
fracture taking place as the load increased.
© 2007 Elsevier B.V. All rights reserved.

Keywords: Nanoscratch; Nanoindenter; Multiscratch; Slip; Scratch friction coefficient

1. Introduction microscope (AFM) the microstructural region to be analysed can


be easily selected and the mechanical properties can be evalu-
Tungsten carbide hard metals are known for their excellent ated on a local scale in the nanometer range [6,7]. Therefore, it is
wear resistance and extensive research has been conducted on possible to measure the individual properties of the binder phase
the macroscopic wear of these materials using such tests as and hard carbide phase and to develop models which describe
the block-on-ring or ball-on-disk as standard testing methods the macroscopic mechanical deformation on the basis of the
[1–3]. However, there is very little understanding available on microscopic properties.
the microscopic mechanical and wear properties of these hard In this paper materials with different grain sizes will be anal-
metals. Earlier work was conducted by Gee et al. to study the ysed and two effects investigated; firstly, the effect of the WC
properties of the constituent phases of WC-Co hard metals by grain size and second, the effect of the cobalt content on the
nanoindentation, but they experienced difficulties with the posi- mechanical and local wear properties of hard metals.
tioning of the stage and a big scatter in the data [4]. Ruff et
al. used a nanoindentation/scratching apparatus with a Vickers 2. Experimental
diamond tip to investigate sliding contact behaviour of ceramic
materials [5]. A similar approach has been adopted in this work 2.1. Materials and sample preparation
to investigate the sliding contact properties of WC-Co hard
metals. With nanoindentation measurements in an atomic force Six commercial hard metal grades with varying cobalt content
ranging from 6 to 15% were tested in this study. The grades
are classified as ultra fine (UFG), medium- and coarse-grained
∗ Corresponding author. Tel.: +49 9131 852 5240; fax: +49 9131 852 7504. depending on the size of the WC grains. The specimens were
E-mail address: [email protected] (S. Ndlovu). prepared for scratch testing by polishing to a 1 ␮m diamond

0043-1648/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2006.11.044
S. Ndlovu et al. / Wear 263 (2007) 1602–1609 1603

Table 1
Properties of the WC-Co hard metals
J15 NY15 T06M T06MF T06MG T06SMG

wt% Co 15 15 6 6.5 6 6
WC grain size (␮m) 2.65 0.25 1.21 0.60 0.48 0.25
Vickers hardness [HV30] 1017 1486 1398 1635 1630 1879

finish. After each polishing step the samples were cleaned in the hardness and modulus of single indents is evaluated as a
ethanol in an ultrasonic bath. The grain sizes and compositions function of displacement.
of all the grades are listed in Table 1.
The material characterisation data on all six grades is pre- 2.4. Instrumented scratch testing
sented and scratch test results on three of the hard metal grades
is discussed. Nanoscratch tests on the hard metals were performed using
a Nanoindenter XP with a load-controlled head. The load is
2.2. Surface morphology of the materials applied normal to the sample surface via a magnet/coil system
which allows for precise and fast control. The indenter column is
The microstructure of the hard metal samples was examined supported by two leaf springs, providing very low stiffness to the
using both an AFM and scanning electron microscope (SEM). A vertical axis. A maximum distance of 1.5 mm is allowed for the
Dimension 3100 atomic force microscope from Veeco, in contact indenter travel normal to the sample surface. Within this range
mode, was used to analyse the scratched samples and the sec- the resolution is less than 0.1 nm. The maximum load capacity
tion analysis function was utilised to examine the scratch profile for the standard system is 500 mN with a precision of less than
allowing for the scratch depth and width to be determined. The 1 mN.
tip frequency and velocity were kept below 1 Hz and 40 ␮m/s, The scratch tests were performed with one corner of the
respectively. A Hitachi S4800 Field-Emission SEM was used to Berkovich indenter in the scratch direction; this was important
study the specimens before and after scratch testing. An accel- as the tip orientation has been found to have an effect on the
erating voltage of 10.0 kV or 15.0 kV was used and the work results [8,9].
distance used was between 15 and 6 mm. The SEM was oper- In this study the normal load ranged from 5 to 400 mN, with
ated in secondary electron imaging (SEI) mode. The WC grain a tip velocity of 0.1 ␮m/s. Tip calibration was performed using
size was determined using the line intercept method and the fused silica before each series of tests to monitor the tip function
Vickers hardness of the materials was measured using a Leco and no significant change in the tip shape was observed.
V-100A macrohardness tester. The scratch samples were examined using an AFM and FE-
SEM and the scratch behaviour of the materials tested was
2.3. Nanoindentation characterised in terms of the scratch width and depth.

For studying the local and global mechanical properties of 3. Results and discussion
the WC-Co hard metals nanoindentation experiments were per-
formed in the low- and high-load regime, respectively. From 3.1. Microstructure
the nanoindentation measurements the hardness and Young’s
modulus of the bulk materials and constituent phase could be To understand the local wear properties of the hard metals, the
determined. microstructure and indentation response of the materials were
The local properties were measured using a nanoindenting studied. Fig. 1 shows the characteristic microstructures of the
AFM (NI-AFM) from Hysitron, Inc., combined with a multi- polished samples. The hard metals consist of angular WC grains
mode AFM from Digital Instruments. This was used to indent with a wide size distribution as indicated in Fig. 2. There are two
the WC grains and cobalt binder phases separately to measure grades containing 15 wt% Co, J15 and NY15 with vary in the
the mechanical properties of each phase in the hard metal. The WC grain size. J15 has a coarse-grained structure with WC grain
indents were carried out with a load of 5 mN. size of 2.65 ␮m and the NY15 is ultra fine grained with a grain
The global properties were investigated at a higher load of size of 0.25 ␮m. The remaining four samples contain 6–6.5 wt%
700 mN with a Nanoindenter XP. Tip shape calibration was car- Co and have a varying WC grain size from 0.25 to 1.2 ␮m. The
ried out using fused silica as the calibration standard. A 4 × 4 T06 samples exhibit high contiguity of the hard metal skele-
array of indents, spaced 20 ␮m apart, were performed on each ton so that some of the WC grains appear very large. In some
sample with a Berkovich indenter using the continuous stiff- cases it was not possible to distinguish the grain boundaries
ness mode of the Nanoindenter XP with a penetration depth clearly.
of 2 ␮m. The Oliver/Pharr procedure is incorporated into the Fig. 3a shows the typical load displacement curves for the dif-
MTS testworks software and was used to evaluate the continu- ferent hard metal samples with the samples containing 15 wt%
ous stiffness measurements (CSM). With CSM measurements cobalt exhibiting the highest indentation depth, whereas the
1604 S. Ndlovu et al. / Wear 263 (2007) 1602–1609

Fig. 1. Characteristic SEM micrographs of the polished (a) WC-6Co and (b) WC-15Co specimens.

lowest penetration depth is exhibited by the 6 wt% Co UFG


sample (T06SMG). The indentation hardness was evaluated at
an indentation depth of around 2 ␮m. In the hardness data two
effects were observed. Firstly, there is a clear dependence of the
hardness on the WC grain size, with a smaller grain size leading
to a higher hardness. Second, a much higher hardness is found
for samples with a lower cobalt content. Young’s modulus was
found to lie between 550 and 750 GPa and in the 6 and 15 wt%
Co samples a clear dependence of modulus on the cobalt con-
tent was observed, with a increasing cobalt content leading to
a decrease in modulus. A similar relationship was observed by
Okomoto et al. [10]. No clear correlation between WC grain size
and Young’s modulus could be established.
The AFM micrographs in Fig. 4a and b show the indents that
were performed on the different hard metals with the same load
Fig. 2. Grain size variation in the hard metal grades studied. of 700 mN, whereas in Fig. 4c the individual WC grains were

Fig. 3. (a) Load displacement curves for the hard metals, (b) indentation hardness vs. WC grain size and (c) Young’s modulus as a function of cobalt weight fraction.
S. Ndlovu et al. / Wear 263 (2007) 1602–1609 1605

Fig. 4. AFM images of indents on (a) J15, (b) T06MF and (C) T06M, (d) SEM micrograph of indent on T06MF showing cracks and glide lines (the arrow indicates
the glide lines).

indented with a load of 5 mN. The SEM micrograph shows more deformed volume under the indenter was calculated. The plas-
details on the deformation that occurred with cracks and glide tically deformed volume is given by
lines visible. The J15 sample containing 15 wt% cobalt exhibits
2
cobalt lips and the shape of the indent observed is typical for π(3ac )3 and 3ac ≈ 8.8hf
sink-in behaviour (Fig. 4a). Sinking in is expected in materials 3
which exhibit a low value of E/Y (where E is the Young’s modu- where ac is the contact radius and hf the maximum depth.
lus and Y is the yield strength); in such materials the plastic zone It was assumed that the WC grains were cuboidal in shape
is normally contained within the boundary of the circle of con- and, therefore, the number of deformed grains was estimated by
tact and the elastic deformations occur further from the indenter. dividing the total deformed volume by the volume of a single
T06MF with a finer microstructure exhibits pile-up behaviour grain.
as shown in Fig. 4b, which is a result of the deflection of WC J15 had the biggest deformed volume but the least number
grains which are pushed upwards by the indenter. Examination of deformed grains because of the coarse-grained structure. The
with a FE-SEM showed that the WC grains at the bottom of the UFG materials contained smaller plastically deformed volumes,
indent contained slip lines so slip was found to be the mecha- but due to the finer microstructure the estimated number of
nism for plastic deformation, and cracks were also observed in deformed grains was significantly higher (see Table 2). There-
some of the WC grains. Slip in WC grains by indentation was fore, the deformation behaviour of the UFG materials is a
also observed by both Engqvist and Jia et al. during indentation bulk material behaviour throughout the entire indentation pro-
with a Vickers indenter, respectively [11,12]. cess, whereas in the coarse-grained materials the individual WC
It was found that the scatter in the hardness decreased at high grains influence the deformation behaviour when the tip first
indentation depths. Moreover, we find a much smaller scatter in enters the hard metal surface.
the hardness for the UFG materials in comparison to the coarser The NI-AFM was used to perform indents on WC grains
grained grades. This is related to the grain size of the hard metals and cobalt binder separately and thereby determine the hard-
and the deformed volume during indentation. To understand the ness and Young’s modulus of the constituent phases of the hard
deformation behaviour of the different hard metals the plastically metal. Due to the low wt% of the cobalt binder phase in the
1606 S. Ndlovu et al. / Wear 263 (2007) 1602–1609

Fig. 5. Scratch friction coefficient graphs for (a) UFG and (b) coarse-grained WC-15Co hard metals.

Table 2 3.2. Scratch testing behaviour


Plastically deformed volume for the hard metals from nanoindentation
Sample WC grain Plastically deformed Number of deformed 3.2.1. Coefficient of friction
size (␮m) volume (␮m3 ) WC grains The scratch tests were conducted at different loads with a
J15 2.65 3510 188 constant tip velocity of 0.1 ␮m/s and at all loads the scratch
NY15 0.25 1591 101845 friction coefficient increased during the first 5–10 ␮m due to
T06MF 0.60 1302 6029 elastic take-up in the instrument before reaching a steady state
T06SMG 0.25 972 62237 of approximately 0.4 (Fig. 5). The scratch friction coefficient
was observed to fluctuate slightly around this value, a result
of the hard WC grains and cobalt phases, which have different
T06 samples and their high contiguity, measurements in the friction behaviour.
binder phase were limited to the J15 sample which contained Subsequently, the fluctuation of the friction coefficient in the
15 wt% cobalt and had a coarse-grained microstructure which UFG materials was lower than that of the coarse-grained materi-
could be easily resolved with the Berkovich tip of the NI-AFM. als. The scratch friction coefficient remained constant when the
Additionally, measurements were not conducted on the fine and scratch load was changed.
ultra fine samples because the poor resolution achieved with
the NI-AFM tip did not allow the microstructure to be clearly
resolved. 3.2.2. Scratch mechanism
The hardness and Young’s modulus of the WC grains was Several damage mechanisms were observed; these were
found to vary in the different samples tested. The WC grains in found to depend on the test conditions and the material
T06M had an average hardness of 25.1 GPa compared with 17.4 microstructure. The overall plastic deformation was a combi-
and 16.7 GPa in J15 and T06MF, respectively. In the sintered nation of brittle and ductile mechanisms.
alloy the carbide particles are randomly orientated so the bulk The coarse-grained WC-15 wt% Co sample (J15) exhibited
alloys exhibit no anisotropy; however, individual tungsten car- grain cracking at a load of 5 mN with cracks forming normal
bide crystals are themselves anisotropic, which would explain to the scratch direction. Multiscratch tests were conducted by
the variation in the results. The cobalt binder in J15 was indented repeated scratching in one direction, with twenty scratches being
and the hardness was found to be 9.4 GPa and Young’s modulus carried out. At 5 mN multiscratching resulted in the chipping
was 342 GPa. The average mean free path in this sample was of WC grains. Some of the WC fragments formed were re-
1.28 ␮m, but this varied so that even during indentation in the embedded in the cobalt binder, as shown in Fig. 6a. So, even
binder phase it was possible to encounter a carbide particle as though the material was damaged the material loss was limited
the tip moved into the surface. by the reintergration of the WC fragments into the cobalt binder.

Fig. 6. SEM images of (a) a multiscratch on J15 at 5 mN and (b) a single scratch at 400 mN.
S. Ndlovu et al. / Wear 263 (2007) 1602–1609 1607

Fig. 7. SEM images of a multiscratch on (a) T06MF and (b) T06SMG at 10 mN load.

Pile-up behaviour typical of ductile metals was limited to the orientation of the WC grains [11]. No grain fall out was observed
ductile cobalt phase. The grain cracking and chipping became in this specimen even at higher loads.
more severe as the load was increased. In the WC-6.5 wt% Co (T06MF) sample similar damage
Examination of J15 at higher loads revealed slip lines on the mechanisms were also observed. However, in addition to crack-
WC grains and in some of the very large grains three different ing and chipping of the WC grains, the fall-out of whole WC
slips planes were observed, as shown in Fig. 6b. Similar observa- grains was also observed with multiscratching at low loads
tions were made by Hegeman and De Hosson after the grinding of 10 mN (Fig. 7) and with single scratches at loads from
of cobalt tungsten carbide with a diamond wheel [13]. The WC 100 mN onwards. Cobalt was smeared on the surface of the
is thought to be deformed by {1 0 1 0} prism slip. Engqvist et UFG 6 wt% cobalt sample (T06SMG) but the scratch friction
al. also observed crack formation from the scratching of WC in coefficient remained unchanged and this sample exhibited less
different crystallographic directions and the direction of the slip WC grain pull out in comparison the medium-grained T06MF
lines formed was found to be dependent on the crystallographic specimen.

Fig. 8. SEM micrographs of the cross-section of a multiscratch on T06MF at 500 mN load.

Fig. 9. A scratch groove on T06MF after scratching with 200 mN load imaged with (a) SEM and (b) AFM; the dotted lines mark the same position in both micrographs.
1608 S. Ndlovu et al. / Wear 263 (2007) 1602–1609

Multiscratching at low loads resulted in WC grain fracture,


chipping and fall out in both 6 wt% samples. At higher loads
increased pulverisation and fragmentation of the WC grains was
observed in addition to more grain fall out. Some of the WC
grains in T06MF also exhibited slip lines, after multiscratching.
No slips lines were observed in the UFG sample; however, the
smearing of cobalt on the surface made it difficult to study the
subsurface grains.
Cross-sections of the scratches on T06MF revealed that there
was no bulk material pile-up during scratch testing in the hard
metal samples studied. The wear groove contained layers of
cracked WC grains, combined with cobalt binder.
During multiscratching there was a gradual build up of
deformed surface layer composed of WC fragments combined
with cobalt binder. This layer covered the length of the scratch
and sections of it were broken off during testing, forming debri
which accumulated on the material surface as indicated in Fig. 8.
The damage mechanisms that take place during scratch test-
ing of WC-Co hard metals are similar to the mechanisms that
have been described during dry sliding wear. Engqvist et al. also
observed the accumulation of wear debris on the surface after
testing of WC-Co hard metals with silicon nitride balls and a
friction coefficient of between 0.3 and 0.5 was measured [14],
in agreement with the results obtained in this work. Pirso et al
also made similar observations after the performance of block-
on-ring tests on WC-Co hard metals and they also observed the
fall out of WC grains after an initial testing period [2].

3.2.3. Scratch profile analysis


Fig. 9 shows the typical appearance of a scratch groove on
WC-Co hard metal when imaged with a FE-SEM and AFM.
The cross-section of the scratches were analysed using the sec-
tion analysis mode of the AFM, thus the scratch depth could be
determined. The scratch width was defined by the extent of the
damage zone. The damage zone was considered to be the area
along the scratch length which had been plastically deformed
by scratching, either by cobalt extrusion, slips lines in the WC
grains, WC grain fall out or cracks in the WC grains. Thus the
scratch width was found to be variable and non-uniform. The
scratch width could be determined from the SEM images as
well as from the AFM analysis and the two values were found
to be different, with the AFM giving consistently slightly larger
values. Fig. 10. Graphs showing the effect of scratch load on (a) scratch width and (b)
The scratch width and depth were found to increase with scratch depth for the WC-Co hard metals and (c) relative increase in scratch
increasing load for single scratches for all the specimens depth as a function of load after 20 multiscratches.
tested (Fig. 10a). As can be seen from Fig. 10b the scratch
depth increased significantly after multiple scratches, a simi- not able to penetrate the surface as effectively, resulting in the
lar response was observed for the scratch width. The increase lower scratch depth and width. Therefore, single scratches at
in the scratch depth after twenty scratches relative to the depth low load resulted in very little damage to the surface structure in
after a single scratch was plotted against the scratch load for comparison to the medium grained T06MF sample which exhib-
the T06MF and T06SMG samples (see Fig. 10c). From this plot ited grain chipping. Thus, the relative increase in scratch depth
it is clear that the two materials respond differently to scratch for the UFG material at low loads was larger because of the
testing. minimal damage that occurred with single scratches. At higher
The scratch depth and width in the UFG 6 wt% Co material loads the damage after single scratches was more severe with
were found to be significantly smaller than that in the fine- grain fracture and fall out in both samples leading to a reduc-
grained material (Fig. 10). The UFG hard metal was harder than tion in the relative increase in the scratch depth with increasing
the medium grained counterpart and thus the diamond tip was load.
S. Ndlovu et al. / Wear 263 (2007) 1602–1609 1609

4. Conclusions References

It was shown in this work that the hard metal microstructure [1] I. Topic, H.G. Sockel, P. Wellmann, M. Göken, The influence of microstruc-
is important in wear testing even when testing is confined to the ture on the magnetic properties of WC/Co hard metals, Mater. Sci. Eng. A
423 (2006) 306–312.
micro or nano scale. Different material deformation mechanisms [2] J. Pirso, M. Viljus, S. Letunovitš, Friction and dry sliding wear behaviour
were observed for the different hard metal grades tested. of cermets, Wear 260 (2005) 815–824.
[3] P.H. Shipway, D.G. McCartney, T. Sudaprasert, Sliding wear behaviour of
1. The cobalt content was found to influence the damage mecha- conventional and nanostructured HVOF sprayed WC-Co coatings, Wear
nisms that occurred. With a high cobalt content the dominant 259 (2005) 820–827.
[4] M.G. Gee, B. Roebuck, P. Lindahl, H.-O. Andren, Constituene phase
damage mechanisms are the plastic deformation of the WC nanoindentation of WC/Co and Ti(C,N) hard metals, Wear A209 (1996)
grains via slip and the formation of intergranular cracks 128–136.
which lead to grain fracture. A low cobalt content leads to [5] A.W. Ruff, H. Shin, C.J. Evans, Damage processes resulting from diamond
grain fall out during scratching at high loads and even during tool indentation and scratching in various environments, Wear 181–183
multiscratch tests at low load. (1995) 551–562.
[6] K. Durst, M. Göken, Micromechanical characterisation of the influence of
2. The WC grain size has also been shown to play a role in Rhenium on the mechanical properties in nickel-base superalloys, Mater.
the wear mechanisms that occur during scratch testing. The Sci. Eng. A 387–389 (2004) 312–316.
ultra fine hard metal grades exhibited very limited slip in [7] B. Bhushan, Nano- to microscale wear and mechanical characterisation
comparison to the coarser grained grades and additionally a using scanning probe microscopy, Wear 251 (2001) 1105–1123.
cobalt film was only observed in the ultra fine grade. [8] S.W. Youn, C.G. Kang, Effect of nanoscratch conditions on both defor-
mation behaviour and wet-etching characteristics of silicon (100) surface,
3. The scratch depth and width were found to increase lin- Wear 261 (2006) 328–337.
early with load and scratch velocity and load were found [9] D. Mulliah, D. Christopher, S.D. Kenny, R. Smith, Nanoscratching of silver
to have no significant effect on the scratch friction coeff- (100) with a diamond tip, Nucl. Instrum. Methods Phys. Res. B202 (2003)
icient. 294–299.
4. A smaller WC grain size resulted in a lower scratch width [10] S. Okomoto, Y. Nakazono, K. Otsuka, Y. Shimoitani, J. Takada, Mechanical
properties of WC/Co cemented carbide with larger WC grain size, Wear
and depth; therefore, better scratch resistance. 55 (2005) 281–287.
[11] H. Engqvist, S. Ederyd, N. Axén, S. Hogmark, Wear 230 (1999) 165–174.
Acknowledgements [12] K. Jia, T.E. Fischer, Wear 200 (1996) 206–214.
[13] J.B.J.W. Hegeman, J.Th.M. De Hosson, Grinding of WC-Co hard metals,
The authors wish to thank Tigra Hartstoff, GmbH for the Wear 248 (2001) 186–196.
[14] H. Engqvist, H. Högberg, G.A. Botton, N. Axén, S. Ederyd, S. Hogmark,
supply of materials. The German Academic Exchange Service Tribofilm formation on cemented carbides in dry sliding conformal contact,
(DAAD) is gratefully acknowledged for its financial support. Wear 239 (2000) 219–228.

You might also like