Sunteți pe pagina 1din 13

Wear 276277 (2012) 1628

Contents lists available at SciVerse ScienceDirect


Wear
j our nal home page: www. el sevi er . com/ l ocat e/ wear
Effect of abrasive particle size and the inuence of microstructure on the wear
mechanisms in wear-resistant materials
M.R. Thakare
a,
, J.A. Wharton
a
, R.J.K. Wood
a
, C. Menger
b
a
National Centre for Advanced Tribology at Southampton (nCATS), Engineering Sciences, University of Southampton, Higheld, Southampton SO17 1BJ, UK
b
Schlumberger Stonehouse Technology Centre, Brunel Way, Stroudwater Business Park, Stonehouse, Glos GL10 3SX, UK
a r t i c l e i n f o
Article history:
Received 27 June 2011
Received in revised form
27 November 2011
Accepted 28 November 2011
Available online 6 December 2011
Keywords:
Wear
WC-based coatings
Deformation
Hardmetals
Size effect
a b s t r a c t
Downhole drilling operations expose tungsten carbide based sintered (WC5.7Co0.3Cr) and sprayed
(WC10Co4Cr) hardmetals to abrasives of different sizes. Although the effect of abradant size on the
abrasive wear of metals has been widely studied, the effect of particle size on the abrasive wear of sin-
tered and sprayed tungsten carbide-based hardmetals has not been examined previously. The abrasion
of hardmetal composite surfaces is complex due to the presence of hard and soft phases which respond
differently during abrasive wear, where an increase in abrasive size leads to a change in the wear mech-
anism which signicantly affects the overall wear rates. Three different abrasive sizes, 4.5 m, 17.5 m
and 180 m, were used in a modied ASTM G65 rubber wheel abrasion test to examine the effects of
abrasive size on wear in a sintered WC and a D-gun sprayed WC-based coating. Uniquely, inuential
parameters affecting the wear mechanisms have been examined and identied with the fundamental
material properties of both abrasives and the multi-phase materials. As a unique way of mapping abra-
sion performance, a parameter previously developed for the micro-abrasion tester, severity of contact,
has been reworked and plotted against a brittleness factor parameter developed in this work. Plotting
these parameters can explain the sharp rise in wear rates associated with the transition from ductile,
plastic deformation dominated material removal to a more fracture-related material removal as the size
of abrasives increases. This work has developed new insights into how hardmetal composites respond to
change in abrasive size and provides a basis for controlling the abrasive particle size.
2011 Elsevier B.V. All rights reserved.
1. Introduction
Downhole drilling operations expose drill-tool components
to extremely aggressive tribological environments. Given their
high hardness and superior wear resistance, WC-based sintered
hardmetals and sprayed coatings are typically used for various
components exposed to these harsh environments. During oper-
ations, the tungsten carbide (WC)-based sintered hardmetals and
sprayed coatings are exposed to a wide size distribution of abrasive
particles, up to 500m, suspended in an alkaline drilling uid (pH
911) [1]. The inuence of abrasive size onwear of ductile materials
has been extensively studied. Typically, abrasive particles respon-
sible for most abrasive wear are between 5mand 500min size,
althoughtheprocess of gougingwear may involveparticles of much
larger size [2]. Misra and Finnie [3] critically reviewed the different
theories explaining the effect of abradants size on the wear rates.

Corresponding author. Present address: Osney Thermouids Research Labora-


tory, Department of Engineering Science, University of Oxford, Oxford OX1 3PJ, UK.
Tel.: +44 1865288760.
E-mail address: mandar.thakare@gmail.com(M.R. Thakare).
The general trend of the specic wear rates (SWRs) observed with
increasing abrasive size for ductile materials is shown in Fig. 1.
Comparing the results fromerosion, grooving and rolling mode
of abrasion, it is reported that for silicon carbide (SiC) particles
smaller than about 100m the wear rates decreased markedly
withdecreasing particle size. Above 100mparticle size, there was
a steady increase in wear rates until about 250m. Stachowiak
and Batchelor [4] found similar results when studying the effect
of increasing abrasive size on the wear rates for ductile materi-
als such as AISI 1096 steel and commercial pure nickel (Ni). One
of the earliest explanations for the size effect is the hard (debris)
layer model, rst proposed by Kramer and Demer [5], suggests that
a surface layer of 50100m thickness becomes work hardened.
Thus, increased abrasion resistance is offered to small particles
or abradants due to the debris layer compared to larger particles
which are able to deform the material below the debris layer and
also abrade material due to plastic deformation. It was further pro-
posed that the work hardening of the subsurface restricted plastic
owandgreater energywas requiredtodeformthis workhardened
subsurface [6]. Moore and Douthwaite [7] proposed the theory of
straindistributionnear the surface toexplainthe creationof a hard-
ened layer of around 10m at the surface. However, the models
0043-1648/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.wear.2011.11.008
M.R. Thakare et al. / Wear 276277 (2012) 1628 17
Fig. 1. Typical wear rate trends observed with respect to abrasive size for ductile
materials [3,4].
discussed in the literature can only be applied for ductile materials
andnot for compositematerials withahardbrittlephaseembedded
in a soft ductile binder.
The wear mechanisms for WC-based hardmetals as well as
sprayed coatings is very different from that of single-phase duc-
tile materials which involves equal pressure of phases or equal
wear of phases as observed by Axen and Jacobson [8]. Examina-
tion of the size effect in AlSiC composites by Candan et al. [9] also
revealed a complex relationship between the SiC grain size within
the composite and the size of the Al
2
O
3
abrasive particles used.
Micrographs of wear scars frommicro-abrasion[10] and ASTMG65
[11] studies on WC-based sintered hardmetals and sprayed coat-
ings conrmthat thewear of suchsurfaces is complicatedduetothe
presence of hard and soft phases within the microstructure which
respond differently during abrasive wear. The response to abrasive
wear may be further inuenced by the pH of the environment and
local variations in pH due to the topography of the wetted surface
and its evolution and the corrosion performance of the WC-based
hardmetals.
The complexity of evaluating the size effect of abrasives arises
from the difculty of entraining abrasives of different sizes as
most wear-tests are designed to entrain abrasives of a spe-
cic size. Comparison of the wear mechanisms observed in the
micro-abrasion tester (using 4.5mabrasives) and ASTMG65 rub-
ber/steel wheel test (abrasive size of 180220m) can provide
greater insights into the wear mechanisms that are likely to occur
when the abrasive particle size changes. However, this does not
give a true reection of the size effect due to the inherent dif-
ferences in the contact conditions and hence cannot be used to
compare wear rates. To overcome these difculties, it is neces-
sary to adapt/modify a well-characterised test set-up to conduct
wear tests using abrasives of different sizes. The present work
investigates the size effect of abrasive particles on the abrasion-
corrosion of a WC-based sintered WC5.7Co0.3Cr and D-gun
(detonation gun) sprayed WC10Co4Cr coating using a modi-
ed ASTM G65 tester by successfully entraining 4.5m, 17.5m
and 180m SiC abrasives whilst keeping all other test conditions
unchanged. Both these materials are extensively used in the oil
and gas industry and hence were selected for the purpose of this
study.
Table 1
Test conditions used during macro-abrasion testing.
Load 20N
Fluid feed NaOH solution of pH 11 or distilled water.
Abrasives SiC abrasives
Abrasive size/feed 180m(fed via hopper), 17.5m(fed via
hopper), 4.5m(pre-mixed slurry)
Abrasive feed rate 120gmin
1
Slurry feed rate 180cm
3
min
1
Slurry pH 11
Counterface material Chlorobutyl rubber (Shore hardness A-60)
Wheel diameter 228mm
Speed of rotation 0.7ms
1
Sliding distances 942m
2. Experimental
2.1. Experimental details
Fig. 2 shows the schematic of the modied ASTMG65 test sys-
tem, consisting of a specimen pressed against a Chlorobutyl rubber
(Shore hardness A-60) rimmed steel wheel of 220mm diameter
under dead-weight loading. The thickness andthe widthof the rub-
ber rim was 12.5mm. A constant load of 20N was applied for the
duration of the test. The sliding distance selected for the test was
942m. Abrasives were fed onto the rubber wheel from a hopper
using a slotted drum mechanism. The drum rotation speed was
maintained such that abrasives were fed at a constant feed rate of
120gmin
1
. To replicate downhole conditions, the tests were car-
ried out under wet conditions using a NaOH solution of pH 11 at
roomtemperature.
A uid feed located in front of the abrasive feed with the nozzle
placed close to the test wheel was used to deliver NaOHsolution of
pH 11 directly on the rubber wheel, see Fig. 2. Due to the inconsis-
tent ow of 4.5m SiC abrasives under dry conditions, they were
pre-mixed in the liquid slurry and fed along with the liquid feed
instead of the hopper-slotted drum mechanism. The volume con-
centration of abrasives in the pre-mixed slurry was maintained
such that it matched the solids to liquid ratio in the experiments
using the larger abrasive sizes.
The overall test matrix is detailed in Table 1. The test speci-
mens (40mm20mm5mm) were ground using a resin bonded
grinding wheel and nally polished using 6mand 1mdiamond
pastes. The resultant surface roughness was 0.01m.
The process of grinding is likely to generate residual stress in
WC specimens, however, some residual stress relieving is achieved
by the nal polishing procedure on the ground specimens [12].
Prior to testing the specimens were cleaned with acetone, dried
and weighed using a Mettler XS205DU precision balance (with a
range of 220g and an accuracy of 0.02mg). This procedure was
repeated after testing and the mass loss was for each specimen was
again measured. Each measurement was repeated ve times. The
specic wear rate SWR (m
3
N
1
m
1
) was calculated by converting
mass loss to volume loss (for known density) and dividing by the
sliding distance and applied load.
2.2. Abrasive size and shape
SEMof the three different abrasives used for the test are shown
in Fig. 4. The abrasive particles were examined under the SEM
and characterised on the basis of their size and shape in order
to identify parameters that inuence their abrasivity. Hamblin
and Stachowiak [13] dened particle abrasivity using a parame-
ter, spike value which was derived by tting a triangle around the
particle edge and measuring the apex angle () and the height of
the triangle (h) as shown in Fig. 3. The spike value was the product
of h and the cosine of the half apex angle (/2). The abrasivity of the
18 M.R. Thakare et al. / Wear 276277 (2012) 1628
Fig. 2. Modied ASTMG65 test rig, NPL.
Fig. 3. Method used to calculate the spike value and the abrasive particle apex angle ().
Fig. 4. Comparison of the abrasives used for the macro-abrasion tests on the modied ASTMG65 rig: (a) 4.5m, (b) 17.5mand (c) 180mabrasive.
M.R. Thakare et al. / Wear 276277 (2012) 1628 19
Table 2
SiC abrasive properties.
Property 4.5m 17.5m 180m
Hardness (GPa) (Hv) 28 (2800)
Elastic Modulus (GPa) 410
Compressive strength (GPa) 2.8
Abrasive particle apex angle in degrees () 6516 8630 12420
Sharpness spike value (Sv) 1.420.35 4.821.91 8.975
Abrasive particle tip radius (m) 1.690.35 6.671.35 18.048.71
Fig. 5. SEMmicrographs showing (a) SEM(BEI) image of a polished surface of WC10Co4Cr and (b) SEM(BEI) image of the polished cross-section of WC10Co4Cr coating.
particles is expected to increase with the spike value, i.e. abrasives
with a higher spike value are likely to be more effective than the
ones with a lower spike value.
The spike value parameter not just considers the sharpness of
the abrasive particle, which is dened by the cosine function of
the half angle, but also the overall size of the abrasive, which is
dened by the height of the triangle. The height of the triangle
can also be interpreted as the tip of the radius of the contact for
each abrasive. For the purpose of this study, random points along
the periphery of the abrasive particles were chosen for calculating
the spike value. The values reported in Table 2 are average values
for all the data thus collected. Whilst the 4.5mabrasive was the
sharpest with an average apex angle of 65

, the 180m SiC abra-


sive had the highest spike value. This clearly makes the 180m
SiC size as the most abrasive amongst those used in the present
work.
2.3. Specimen microstructure
Fig. 5a shows the SEM-backscattered electron image (BEI)
image of a polishedD-gun(detonationgun) sprayedWC10Co4Cr
surface. The darker regions in the BEI image represent the heav-
ier elements, such as tungsten, in the composition and the
lighter regions represent the lighter elements, such as cobalt and
chromiumin the composition. The micrograph shows an inhomo-
geneous distribution of carbide rich and binder rich areas along
with the presence of some voids and cracks formed on the surface
during the cooling of the coating. The inhomogeneous structure of
the coating is also reected in the larger scatter (100Hv) seen
for the hardness measurements of the sprayed coatings compared
to the smaller scatter (45Hv) seen for sintered specimens, see
Table 3.
The SEM micrographs also reveal the non-uniform size distri-
bution of the carbide grains in the coatings. The coating appears
to have larger carbide (34m) grains towards the centre of the
splats andner carbidegrains, less than1minsize, alongthesplat
boundaries. This is likely to occur because of the higher degree of
decarburising of WC grains along the splat boundary as the temper-
ature along the splat boundary is expected to be relatively higher
in comparison with the centre of the splat [18]. The SEM of a pol-
ished cross-section of WC10Co4Cr coating is shown in Fig. 5b.
The cross-section also shows the presence of carbide and binder
rich areas in the coating along with a possible splat boundary. In
general, the splats were approximately 50m wide and 10m
thick. Earlier work [19] on this coating had revealed the presence
of W
2
C and metallic Wring around the carbide grains in the coat-
ing formed as a result of decarburisation during the spray process
and the subsequent re-precipitation during cooling.Fig. 6a shows a
SEMmicrograph of a polished surface of sintered WC5.7Co0.3Cr
specimen, revealing the skeletal carbide structure, typical of sin-
tered hardmetals. The size of the carbides was 23m. The skeletal
structure was also observed in the cross-section (Fig. 6b). Unlike
the WC10Co4Cr coating, the sintered specimens revealed a rel-
atively homogeneous microstructure with no evidence of defects
Table 3
Mechanical properties and designations of test specimen.
Composition Hardness (Hv)
(GPa)
Youngs
modulus (E)
(GPa)
Fracture
toughness (K
1C
)
(GPa

m)
Carbide
size (m)
Coating
thickness
(m)
Coating
type
Substrate Density
a
(gcm
3
)
Source
WC5.7Co0.3Cr 177345
(17.39)
600 [14] 710
3
[15] 2.3 15.28 Dymet Alloys,
UK
WC10Co4Cr 1114100
(10.92)
300 [16] 410
3
[17] 24 250300 Detonation
gun
UNS S316 14.68 Praxair Surface
Technologies
a
Densities of specimens calculated frommaterial composition.
20 M.R. Thakare et al. / Wear 276277 (2012) 1628
Fig. 6. Polished surface of a sintered WC5.7Co0.3Cr specimen showing the typical skeletal structure of the carbides.
such as voids and cracks. WC-based sintered composites are aggre-
gates of particles of tungsten carbide bonded with a binder phase
by liquid-phase sintering [20,21]. During the sintering of WC-based
hardmetals, treated powder of WC and binder is heated to approxi-
mately 1500

Cwithanappliedpressure of about 70100MPa [22].


During sintering, substantial amounts of WC dissolve in Co and re-
precipitate during cooling, mainly along the periphery of existing
WC grains and as nely dispersed particles in the binder which
results in a strong bond between the carbide and the Co binder.
This is due to the ability of Co to dissolve WC at high temperatures
(up to 35%) [22].
3. Results
3.1. Abrasive size effect in WC10Co4Cr coating
Fig. 7 shows the specic wear rates (SWR) for sprayed
WC10Co4Cr coating specimens abraded using 4.5m, 17.5m
and 180m SiC abrasives. The wear rates for the 4.5m and
17.5m abrasives are very similar but doubles in value for the
180mabrasive. Toelucidate the factors inuencingthe wear rates
and to determine the predominant wear mechanisms, SEMmicro-
graphs of the worn surfaces were examined. For consistency, the
SEManalysis was performed on the centre of the wear scar for all
specimens. Transverse-sections werealsopreparedacross thewear
scar for select specimens for further analysis.
Fig. 7. Specic wear rates for WC10Co4Cr coating during wet rubber wheel abra-
sion test (pH 11) using 4.5m, 17.5mand 180mSiC abrasives.
Fig. 8 compares the SEMmicrographs of wear scars for the three
different abrasive sizes. As seen from Fig. 8a, abrasive wear by
4.5mSiC particles results in a preferential removal of the binder-
phase followed by the undermining and subsequent removal of
unsupported carbide grains. This is conrmed by the presence of
carbide grain cavities in the wear scar. The removal of the binder
phase around the carbide grains appears to be caused by plastic
grooving (two-body abrasion) which would also result in the sub-
sequent ejection of unsupported carbide grains. The preferential
removal of the binder is clearly visible in the high magnication
image of the same specimen, Fig. 8b. In addition to the removal of
the binder-phase by two-body grooving, there appears to be signs
of washing away of the binder, which is a oweffect, possibly due
to the relative motion between the rubber wheel and the specimen
surface and the presence of fragmented abrasive particles (smaller
than 1min size) in the slurry.
Fig. 8c shows the specimen worn using the 17.5mSiC abrasive
has a grooved surface. The grooves appear to be signicantly wider
and there are fewer carbide grains present within them as com-
pared to the specimen worn using the 4.5m abrasive. At higher
magnication (Fig. 8d) a single carbide grain is evident within the
groove withsome preferential binder depletionaroundit, exposing
the carbide grain to abrasive particles. Interestingly, the number of
extant carbide grains within the grooves appears to be lower than
those seen in the specimen worn using the 4.5m abrasive. This
indicates that an increase in the size of the abrasive particles (to
10 times the size of average carbide grain) results in carbide par-
ticles offering less resistance to plastic deformation and grooving.
Overall, wear by the 17.5mabrasive appears to be due abrasive-
size plastic grooving through the coating, removing the bulk of
material including carbide grains. Additionally, the motion of the
abrasive particles also results in the fragmentation and removal of
unsupported carbide grains within the wear grooves.
The wear mechanism in the carbide-rich areas on the coating
surface is consistent with the equal pressure on the phases mode
predicted by Axen and Jacobson [8] originally used for understand-
ing wear of sintered hardmetals. Equal pressure on the phases is
a situation in which equal amounts of pressure is borne by the
hard as well as the soft phases within the structure. Consequently,
the soft phase is expected to undergo mores extensive wear com-
pared to the hard phase resulting in the preferential abrasion of the
softer phase (binder). This preferential removal of the soft phase is
expected to expose and over a period of time, undermine the
harder phase (carbides), which can subsequently be abraded. In
case of the coating, the undermining of the hard particles results
in the ejection of loosely held carbides from the surface as can be
inferred fromthe presence of cavities in the wear scar.
M.R. Thakare et al. / Wear 276277 (2012) 1628 21
Fig. 8. SEMmicrographs of wear scars onWC10Co4Cr coated sample (a) wornusing 4.5mabrasives, (b) highmagnicationimage of (a), (c) wornusing 17.5mabrasives,
(d) high magnication image of (c), (e) worn using 180mabrasives and (f) high magnication image of (e) (direction of abrasive motion: right to left).
The SEM of a specimen worn using 180m SiC abrasives is
shown in Fig. 8e. Interestingly, the worn surface shows the pres-
ence of ne grooves (1mwide) and unlike the specimens worn
using 4.5mand 17.5mabrasives, the grooves on the surface are
not similar in size to that of the abrasive used (i.e. 180m). Fur-
thermore, the high SEM magnication of the same area (Fig. 8f),
does not showany signs of extant carbides. Additionally, the wear
scar surface shows presence of transverse cracks in the wear scar
running perpendicular to the direction of the abrasive motion, sug-
gesting sub-surface damage and brittle fracture occurring in the
coating due to high contact stresses caused by the increase in the
size of abrasive particle (100 times the size of the average carbide
grain). The presence of ne grooves on the surface also indicates
that the initial high contact stress leads to fragmentation of the
abrasives along with the brittle fracture of the coating. The absence
of surface cavities inthe wear scar formedby the removal of unsup-
portedcarbides also indicates that the material loss was mostly due
to the propagation of cracks and formation of splat-sized debris.
Tofurther investigatethewear mechanisms, transversesections
through the wear scars of the three specimens are shown in Fig. 9.
Plastic grooving by 4.5mabrasives is also conrmed by the pres-
ence of 45m wide grooves on the surface of the wear-scar as
shown in Fig. 8a. Interestingly, the stresses induced by the contact
of 4.5mabrasives do not appear to cause sub-surface cracking in
the coating. Similar grooves caused by abrasive particles plough-
ing through the coating were also observed for the specimen worn
using 17.5m abrasives, see Fig. 9b. As with the specimen worn
using 4.5m abrasives, the size of the grooves was similar to the
size of the abrasives. However, the cross-section of worn surface
using 17.5m abrasive also shows some signs of lateral cracking
22 M.R. Thakare et al. / Wear 276277 (2012) 1628
Fig. 9. Comparison of cross-sections of WC10Co4Cr coated specimen worn using: (a) 4.5mabrasives, (b) 17.5mabrasives and (c) 180mabrasives.
in the grooves. Brittle fracture of HVOF WCCo coating has been
observed during indentation by hard particles either in the formof
Palmqvist cracks or Median cracks (half penny crack) depending on
the indentation loads [23].
The cross-section of the coating worn using 180m abrasive
particles reveals a dramatic increase in the levels of sub-surface
cracking. However, unlike the cross-sections of coating specimens
worn using 4.5m and 17.5m abrasives, the size of the grooves
are not similar to the size of the abrasives used. It has been shown
[24] that indentation by hard particles can cause propagation of
cracks that leadto delaminationof the coating insplat-sizeddebris.
The absence of abrasive size grooves also suggests that the initial
entrainment of the 180mabrasives is likely to result in the frag-
mentation of the abrasive particles along with the formation of
sub-surface lateral cracks. The formation and propagation of lat-
eral cracks is consistent with the model of abrasive wear of brittle
surfaces proposed in the literature [2]. The subsequent propaga-
tion of sub-surface lateral cracks within the coating would result in
the removal of splat-size debris. As seen from the SEM study, the
propagation of sub-surface lateral cracks appears to have occurred
through the interface of the outer region of the splat and the inner
core. The interface between the outer regions of the splat and the
inner core is expected to be provide a natural source due to the dis-
solution of carbides to form brittle ternary carbides and WMC
compounds during spraying [25]. HVOF and D-gun-sprayed WC-
based coatings are also known to have lower fracture toughness
parallel to substrate (along upper splat boundaries) compared to
perpendicular [15].
3.2. Sintered WC5.7Co0.3Cr
Fig. 10shows the wear rates for sinteredWC5.7Co0.3Cr speci-
mens abradedusing4.5m, 17.5mand180mSiCabrasives. The
SWR using 4.5m and 17.5m abrasives was found to be similar
whilst the wear rate using 180mabrasives increased by an order
of magnitudethus indicatingasignicant changeinthewear mech-
anism. For better representation, the SWR (y-axis) is presented on
a logarithmic scale
The worn sintered surfaces were examined to determine the
wear mechanisms occurring. For consistency, the SEM investiga-
tionwas undertakenat the centre of the wear scar andthe direction
of abrasive motion was from right to left. Fig. 11a shows the SEM
analysis of sintered WC5.7Co0.3Cr worn using the 4.5m SiC
abrasive and reveals the depletion of the binder-phase around the
extant carbide grains and cavities indicating the removal of unsup-
portedcarbide grains. At higher magnication(Fig. 11b) the carbide
structure appears to be intact with some evidence of carbide grain
cracking. The lack of fragmentation of carbide grains appears to
suggest that the overall wear mechanism is dominated by the
depletion and washout of the binder phase around the carbide
grains leading to the weakening of the carbide structure. Removal
of unsupported carbide grains is likely to be the consequence of
this weakening. Abrasive wear of WC-based sintered hardmetals
using ne abrasives (typically similar in size as the carbide grains)
reveals a two-stage wear mechanism. A similar wear mechanism
was observed by Shipway and Hogg [10] during the micro-abrasion
of WC-based sintered hardmetals.
Fig. 11c shows the wear scar on the sintered specimen worn
using17.5mSiCabrasives. Thewear scar reveals askeletal carbide
structure devoid of binder, the main difference with the surface
worn using the 4.5m abrasive being the fact that most carbide
grains appear to be compacted and pushed against each other.
However, as is the case with the specimen worn by the 4.5m
abrasive, there is very little evidence of fragmentation and loss of
carbide grains within the skeletal structure. High magnication
assessment of the wear scar further conrms the lack of car-
bide grain loss and also shows that the binder-phase is squeezed
between the carbide grains. The mechanism of binder removal is
found to be signicantly different from that observed when using
4.5m abrasives. With the larger 17.5m abrasive particles (10
times the size of carbide grains), the preferential removal of the
binder phase is due to the extrusion of the binder phase between
the carbide grains as predicted by the Larsen-Basse [26] model.
This model considers that the binder is squeezed out of the sur-
face as a result of compressive stresses along the sliding direction
during abrasionas a result of the signicant differences inthe resis-
tance to plastic deformation between the carbide grains and the
binder-phase. An empirical relationship proposed by Larsen-Basse
[26] in this model suggests that the amount of binder extruded was
Fig. 10. Specic wear rates for sintered WC5.7Co0.3Cr during wet rubber wheel
abrasion test (pH 11) using 4.5m, 17.5mand 180mSiC abrasives.
M.R. Thakare et al. / Wear 276277 (2012) 1628 23
Fig. 11. SEMmicrographs of wear scars of sintered WC5.7Co0.3Cr specimens: (a) worn using 4.5mabrasives, (b) high magnication image of (a), (c) worn using 17.5m
abrasives, (d) high magnication image of (c), (e) worn using 180mabrasives and (f) high magnication image of (e) (direction of abrasive motion: right to left).
directly proportional to the size of carbide particles and the total
strain developed within the hardmetal and inversely proportional
to the binder mean free path.
The SEM of the sintered specimen worn using the 180m SiC
abrasive is shown in Fig. 11e. The micrograph reveals a severely
fragmented carbide structure with a very different morphology to
that seen in samples abraded by 4.5m and 17.5m abrasives.
In this case, the overall wear appears to be due to the loss of
binder as well as the fragmentation of individual carbide grains
characterised by absence of binder at the surface as well as the
extensively deformed/fragmented carbide grains. To elucidate the
mechanism further, the wear scar was also examined at higher
magnication, see Fig. 11f. The examination of the high magni-
cation image reveals the depletion of the binder-phase around the
carbide grains and the cracking of individual carbide grains along
their slip planes, with the washing away of unsupported fragments
towards the later stages of the overall wear process. In the ini-
tial stage of abrasive wear, abrasive particles selectively remove
the binder phase leading to the formation of small pits with inter-
granular facets. As reported by Blombery et al. [27], the removal
of the binder-phase leads to the lowering of fracture strength and
development of plastic strain within the surface layers of the sin-
tered hardmetal. The strain developed is relieved by the formation
of cracks within the carbide grains along their slip planes [2729].
The extensive removal of the binder phase andthe subsequent frag-
mentation of the carbide grains is likely to be most effective with
the 180mabrasives which would exert high contact stress during
the initial entrainment resulting in maximumstrain and would aid
binder removal by subsequent fragmentation as predicted by the
Larsen-Basse [26] model.
24 M.R. Thakare et al. / Wear 276277 (2012) 1628
Fig. 12. Comparison of the cross-sections of sintered WC5.7Co0.3Cr specimens worn using (a) 4.5m, (b) 17.5mand (c) 180mabrasives.
Table 4
Summary of wear mechanisms.
Size Sprayed WC10Co4Cr Sintered WC5.7Co0.3Cr
4.5m Preferential removal of the
binder phase by wash-out
Removal of carbides by
undermining
Preferential removal of the
binder-phase Removal of
unsupported carbide grain
fragments
17.5m Removal of binder by
plastic deformation and
grooving
Undermining and
fragmentation of carbide
grains
Removal of binder by
extrusion method
Removal of carbide grain
fragments by undermining
180m Removal of splat-size
debris by formation and
propagation of lateral
cracks in the coating
Severe binder extrusion
Extensive cracking of
carbide grains and removal
of carbide fragments
Fig. 12 shows the cross-sections of the worn specimen abraded
using 4.5m, 17.5m and 180m SiC abrasives. There is no evi-
dence of sub-surface crack propagation in any of the specimens,
suggesting that the damage induced when using the three abra-
sive sizes was local and does not result in large-scale delamination.
This is also an indicator of the better resistance of the skele-
tal microstructure to crack propagation compared to the lamellar
microstructure of HVOF coatings. A possible reason for this is the
presence of compressive stresses in the WC grains and tensile
stresses in the binder-phase induced during the sintering process,
result in WCCo composite not encouraging propagation of cracks
through the bulk material [27].
3.3. Summary of wear mechanisms
For both sintered WC5.7Co0.3Cr and the WC10Co4Cr coat-
ing, there is a clear size effect evident, with a similar wear
mechanism being observed for 4.5m and 17.5m SiC abra-
sives which then changes dramatically when the abrasive size is
increased to 180m. Table 4 summarises the various wear mech-
anisms observed.
To gain a better understanding of the effects of abrasive par-
ticle size on the contact conditions, parameters such as number
of particles in the contact and load per particle was calculated
for each abrasive size for the conditions used during testing,
see Table 5.
4. Discussion
In their study of the abrasive wear of brittle solids, Moore
and King [30] predicted that material removal by fracture
Table 5
Analysis of wear scar and related parameters.
Abrasive size (m) 4.5 17.5 180
Load (N) 20
Wear scar area (mm
2
)
(assuming typical wear
scar of 15mm15mm)
225
Wear scar depth (m) in
terms of depth (m),
carbide worn, splats
worn
1012m, 23
carbides/1
splat
1214m, 34
carbides/1
splat
2527m,
810
carbides/23
splat
Groove size (m) 45 1618 <1
Mass of single abrasive
particle (g)
2.9210
10
1.7210
8
1.8710
5
Number of
particles/contact area
1.4710
8
2.4910
6
2.2910
3
Load per particle (N) 1.3710
7
8.0310
6
8.7410
3
mechanism was likely to result in almost an order of magnitude
increase in the wear rate as compared to material removal by plas-
tic deformation. The factors likely to inuence the transition of the
dominant wear mechanism from plastic deformation to fracture
were:

Increase in depth of indentation caused by abrasive.

Increase in sharpness of abrasive.

Decrease inthe ratio of fracture toughness to hardness of wearing


material.
Lawn et al. [31] proposed that for the transition from plastic
deformation to fracture to occur, the depth of indentation must
exceed the critical indentation size and was proportional to the
ratio (H
2
/E), where H is the hardness, is the fracture surface
energy and E is the Youngs modulus of the material. Similarly, Zok
and Miserez [32] used the ratio of (K
2
C
/H
3
) to predict resistance to
cracking in brittle materials. For the same material, increase in the
critical indentationsize andhence the transitionfromplastic defor-
mation to fracture is likely to occur with either an increase in the
depthof indentationor theincreaseinthesharpness of theabrasive.
Clearly, the indentation depth during abrasive wear is directly pro-
portional to the load per particle and hence the size of the abrasives
used as summarised in Table 5.
A measure of the severity of contact has been developed
by Adachi and Hutchings [33] to evaluate the transition from
grooving to rolling mode during micro-abrasion. Although the
contact geometry during the rubber-wheel wear test is very
different to the micro-abrasion test, the severity of contact can
be considered a useful parameter for comparing the effects of
change in abrasive size on wear rates. Unlike the micro-abrasion
M.R. Thakare et al. / Wear 276277 (2012) 1628 25
Fig. 13. Severity of contact vs. abrasive size. Additional data on bulk materials and coatings plotted fromliterature [3537].
test which has non-conformal point contact geometry to begin
with, the rubber-wheel test results in a relative large contact area
between the rubber and the specimen upon loading. An empirical
formula for calculating the number of particles within the contact
at any given time, proposed by Stevenson and Hutchings [34] for
dry rubber wheel tests, was used to estimate the number of abra-
sives within the contact zone. The formula, see Eq. (1), considers
factors such as mass ow rate past the specimen in gmin
1
(m
f
),
contact length (x), relative velocity () and mass of typical abrasive
particle (m
p
) for calculating the number of particles in the contact
zone. Considering that the size of abrasives used for this study
was considerably smaller than the typical sand grains used for
rubber wheel test used by Stevenson and Hutchings and that the
entrainment of abrasives is likely to be greater in wet conditions,
the mass ow rate of particles past the specimen m
f
was assumed
to be equal to the mass ow rate of the abrasives used in the
test.
N
P
=
m
f
x
vm
p
(1)
As shown in Table 5, the three different sizes result in very
different load per particle values. It seems fairly obvious that an
increase in the size of abrasives would result in an increase in the
load per particle. The severity of contact as originally developed
[33] did not consider the effects of abrasive properties. However,
the severity of the loaded contact is likely to depend on abra-
sive properties such as hardness, shape and fracture toughness.
Hence, the severity of contact expression used in present work was
modied to include the ratio of abrasive hardness and specimen
hardness (H
Abr
/H
WC
). Also, shape and fracture tendency of abra-
sives was included to reect the increase in severity when using
a sharp abrasive with greater fracture toughness. Therefore, the
severity of contact (S
c
) expression for the rubber-wheel test can be
written as:
S
C
=
load
particle
(Sv)
Abr
(K
1C
)
Abr
H
Abr
H
S
(2)
where Sv is the spike factor of the abrasive particle; K
1C
is the frac-
ture toughness of the abrasive material measured frombulk; H
Abr
is the SiC hardness; and H
S
is the hardness of specimen.
As seen from the examination of the worn surfaces, it is clear
that the change in wear mechanism is directly related to the size
of the abrasive particles and the resultant load per particle. To help
with the analysis, a selection of papers fromthe literature using the
rubber wheel abrasion test to abrade a range of material and abra-
sive combinations have beenexaminedfor comparison. This allows
comparison between the severity of contact and SWR for a range
of abrasives (silica sand, SiC and alumina) and a range of materi-
als (steels, ceramics, coatings). The severity of contact calculated
for all the material-abrasive combinations reviewed is graphically
represented in Fig. 13.
A trend of increased load per particle with increasing abrasive
size can be seen for ductile as well as brittle materials. Moreover,
the severity of contact is also affected by the difference in the hard-
ness of the abrasives used and the material being worn, as can be
seenfromthe examples of mildsteel andWC(sinteredandcoating)
being worn using 125mand 225msand abrasives respectively.
Greater hardness for theWCspecimens results inalmost twoorders
26 M.R. Thakare et al. / Wear 276277 (2012) 1628
Fig. 14. SWR vs. severity of contact for various materials tested using the wet rubber wheel test as part of this programme and fromthe literature [3537].
of magnitude decrease in the severity of contact. Consequently, the
wear rates also show a similar trend with the WC samples result-
ing in two orders of magnitude lower wear as compared to the mild
steel samples, see Fig. 14. Clearly, severity of the contact, incorpo-
rating factors such as the ratio of hardness of the abrasives and
sample, fracture toughness of the abrasives, load per particle and
the abrasive concentration will inuence the rate as well as the
mechanism of wear observed. However, as seen from the results,
the overall wear rates and mechanisms are also expected to be con-
trolled by the inherent tendency of the sample to fracture under
increasedindentationloading. Fig. 14compares the wear rates with
severity of contact for all the materials examined in Fig. 13. As
expected, SWR linearly increases with an increase in the severity
of contact. However, more interestingly, these trends can be sep-
arated by abrasive size groups as highlighted in Fig. 14 by straight
lines drawn through data points. In the largest set of results exam-
ined for the abrasive size group 125225m, all samples worn
irrespective of sample hardness, abrasive type, fracture toughness,
hardness and spike factors appear to be stacked in a region indi-
cated using a green box. Another interesting feature highlighting
the effect of abrasive size on the wear rates is the rise in the slope
of lines connecting data points withanincrease inthe abrasive size.
Marshall and Lawn [38] suggested that the ratio of (H/K
C
) of a
material is also a measure of brittleness and together with the ratio
of thematerials hardness andYoungs modulus (H/E), couldbeused
to predict the tendency of a material to fracture. Thus, the product
of (H/K
C
)
2
and (H/E) could be termed as brittleness factor of that
material. Fig. 15 shows a graph plotting the severity of contact vs.
the brittleness factor for all the samples examined in Fig. 13.
Interestingly, the materials which show greater tendency to
fracture wouldhave higher brittleness factor andwouldbe towards
the right of the x-axis, whereas, materials likely to show a more
ductile-type failure would be towards the left of the x-axis. How-
ever, even for the brittle materials to fail by fracture, the severity
of contact must exceed a minimum value. For example, the
WC10Co4Cr coating shows plastic grooving and ductile mode
of failure at 10
6
severity of contact but a mixed mode, with duc-
tile, plastic and some lateral cracking at 10
4
severity of contact.
With a subsequent increase in the severity of contact to 1, the coat-
ing appears to fail purely by lateral cracking. Clearly, the value of
severity of contact at which the failure mode changes from duc-
tile to brittle is likely to increase with a decrease in the brittleness
factor. The dominant wear mechanisms observed in the samples
included in Fig. 14 were identied by triangles of different colours.
Interestingly, the ones that showedplastic deformation-dominated
material removal mechanisms were identied as the samples with
either lower brittleness factor or lower severity of contact (indi-
cated by green triangles). Conversely, the samples that showed
fracture dominated material removal mechanisms were identied
as the ones with higher brittleness factor and/or severity of contact
(indicated by red triangle). Although the current set of data does
not map out the entire region shown in the graph, there appears to
be a general trend, whereby materials with low values of brittle-
ness factor andseveritywouldresult inaductile-typefailure, whilst
materials with high values of brittleness factor and severity of con-
tact were likely to fail froma fracture-dominated mechanism. The
regions are identied by dotted line drawn between the triangles
in Fig. 15.
M.R. Thakare et al. / Wear 276277 (2012) 1628 27
Fig. 15. Severity of contact vs. brittleness factor for various materials tested using the wet rubber wheel test as part of this programme and fromthe literature [3537]. Brittle
failures are stacked towards the top-right corner of the chart whilst ductile, plastic deformation dominated wear towards the bottom-left corner.
5. Conclusions
Unlike single-phase materials, for the WC-based sintered hard-
metal and sprayed coatings an increase in abrasive size leads to a
change in the overall wear mechanism which signicantly affects
the overall wear rates. As a result of the change in wear mecha-
nisms, adistinct size-effect is observedfor bothWC-basedsintered
hardmetal and D-gun sprayed coating examined in the present
work.
This changeinwear mechanismis not just relatedtotheentrain-
ment of abrasives, i.e. change from grooving to rolling abrasion,
but also results in a change in material removal mechanisms from
ductile, plastic deformation to a more fracture-dominated failure
mode. This transition was observed for both sintered hardmetal
as well as the D-gun sprayed coating. For each material, a critical
mechanisms was identiedthat leadto a sharpincrease inthe wear
rates, usually associated with an increase in the abrasive size used.
For WC10Co4Cr coating, theformationandpropagationof lateral
cracks was identied as the step at which wear rates showa sharp
increase, occurring when using 180m SiC abrasives. Similarly,
for sintered WC5.7Co0.3Cr, it was the fragmentation of carbide
grains that led to a sharp rise in the wear rates, which occurred
when using 180mSiC abrasives. Overall, a change in mechanism
as a result of an increase in the abrasive size was attributed to
the increase in the severity of contact associated an increased load
per particle. The severity of contact parameter was redened using
abrasive as well as the wearing material properties. An interest-
ing correlation between wear rate and the severity of contact was
established for various materials which clearly highlighted the size
effect.
In addition to the severity of contact, it can be concluded that
this transition also depended on the parameter classied as the
brittleness factor. Data fromthe literature has also used to support
this theory and demonstrate that increasing the severity of contact,
along with an increase in the brittleness factor, would result in the
transition froma ductile to fracture dominated wear mechanism.
Acknowledgments
The authors would like to acknowledge Schlumberger plc for
funding this work and the management of National Physical Labo-
ratory (NPL) for granting access to the modied ASTMG65 rig. The
authors would also like to express their gratitude to Dr. Andrew
Gant andProfessor MarkGee for their time andinvaluable guidance
during the experimentation at NPL.
References
[1] Minutes from1819 April 2005 visit to Stonehouse Technology Centre, Private
communication, April 2005.
[2] I.M. Hutchings, Tribology: Friction and Wear of Engineering Materials, Edward
Arnold, London, 1992.
[3] A. Misra, I. Finnie, On the size effect in abrasive and erosive wear, Wear 65
(1981) 359373.
[4] G.W. Stachowiak, A. Batchelor, Engineering Tribology, 3rded., Elsevier, London,
2002.
[5] I. Kramer, L. Demer, The effect of surface removal on the plastic behaviour of
aluminiumsingle crystal, Trans. AIME 221 (1961) 780786.
28 M.R. Thakare et al. / Wear 276277 (2012) 1628
[6] H. C imeno glu, Subsurface characteristics of an abraded lowcarbon steel, Wear
210 (12) (1997) 204210.
[7] M. Moore, R. Douthwaite, Plastic deformation below worn surface, Metall.
Trans. (1976) 18331839.
[8] N. Axen, S. Jacobson, A model for abrasive wear resistance for multi-phase
materials, Wear 174 (1994) 187199.
[9] E. Candan, H. Ahlatci, H. C imeno glu, Abrasive wear behaviour of AlSiC com-
posites produced by pressure inltration technique, Wear 247 (2) (2001)
133138.
[10] P. Shipway, J.J. Hogg, Dependence of micro-scale abrasion mechanism of
WCCo hardmetals on abrasive type, Wear 259 (2005) 4451.
[11] A.J. Gant, M.G. Gee, B.R. Roebuck, Rotatingwheel abrasionof WC/Cohardmetals,
Wear 258 (2005) 178188.
[12] Meeting minutes-with Dr. M. Gee and Dr. A. Gant, NPL, 2007.
[13] M.G. Hamblin, G.W. Stachowiak, A multi-scale measure of particle abrasivity,
Wear 185 (1995) 225233.
[14] H. Doi, Y. Fujiwara, K. Miyake, Y. Oosawa, A systematic investigation of elastic
moduli of WCCo alloys, Metall. Trans. 1 (1970) 14171425.
[15] S.F. Wayne, S. Sampath, Structure/property relationships in sintered
and thermally sprayed WCCo, J. Therm. Spray Technol. 1 (4) (1992)
307315.
[16] G. Sundararajan, P.S. Babu, DetonationsprayedWC-Cocoatings: unique aspects
of their structure and mechanical behaviour, Trans. Indian Inst. Metals 62 (2)
(2009) 95103.
[17] J. Barber, B.G. Mellor, R.J.K. Wood, The development of sub-surface damage
during high energy solid particle erosion of a thermally sprayed WCCoCr
coating, Wear 259 (2005) 125134.
[18] D. Stewart, P. Shipway, D.G. McCartney, Microstructural evolution in thermally
sprayed WCCo coatings: comparison between nanocomposite and conven-
tional starting powder, Acta Mater. 40 (2000) 15961604.
[19] M.R. Thakare, J.A. Wharton, R.J.K. Wood, C. Menger, Exposure effects of strong
alkaline conditions on the micro-scale abrasion-corrosion of D-gun sprayed
WC10Co4Cr coating, Tribol. Int. 41 (2008) 629639.
[20] Z. Yao, J. J. Stiglich, T. S. Sudarshan, Nano-grained Tungsten
CarbideCobalt (WC/Co), Materials Modication, Inc., www.matmod.com/
Publications/armor 1.pdf.
[21] W. Dawihl, B. Frisch, Wear properties of tungsten carbide and aluminiumoxide
sintered materials, Wear 12 (1968) 1725.
[22] C.J. Smithells, Tungsten-A Treatise on Its Metallurgy, Properties and Applica-
tions, 3rd ed., Chapman and Hall Ltd, London, 1952.
[23] M.M. Lima, C. Godoy, J.C. Avelar-Batista, P.J. Modenesi, Toughness evaluation
of HVOF WCCo coatings using non-linear regression analysis, Mater. Sci. Eng.
A357 (2003) 337345.
[24] E.L. Cantera, B.G. Mellor, Fracture toughness and crack morphologies in eroded
WCCoCr thermally sprayed coatings, Marer. Lett. 37 (1998) 201210.
[25] R.J.K. Wood, B.G. Mellor, M.L. Bineld, Sand erosion performance of detona-
tiongunappliedtungstencarbide/cobaltchromiumcoatings, Wear 211(1997)
7083.
[26] J. Larsen-Basse, Binder extrusion in sliding wear of WCCo alloys, Wear 105
(1985) 247256.
[27] R.I. Blombery, C.M. Perrot, P.M. Robinson, Abrasive wear of tungsten
carbidecobalt composites. I. Wear mechanisms, Mater. Sci. Eng. 13 (1974)
93100.
[28] M. Gee, C. Phatak, R. Darling, Determination of wear mechanisms by stepwise
erosion and stereological analysis, Wear 258 (14) (2005) 412425.
[29] M. Gee, A. Gant, B. Roebuck, Wear mechanisms in abrasion and erosion of
WC/Co and related hardmetals, Wear 263 (2007) 137148.
[30] M.A. Moore, F.S. King, Abrasive wear of brittle materials, Wear 60 (1980)
123140.
[31] B.R. Lawn, T. Jensen, A. Arora, Brittleness as an indentation size effect, J. Mater.
Sci. 11 (1976) 573575.
[32] F.W. Zok, A. Miserez, Property maps for abrasion resistance of materials, Acta
Mater. 55 (2007) 63656371.
[33] K. Adachi, I. Hutchings, Sensitivity of wear rates in the micro-abrasion test to
test conditions and material hardness, Wear 258 (2005) 318321.
[34] A. Stevenson, I. Hutchings, Development of the dry sand/rubber wheel abrasion
test, Wear 195 (1996) 232240.
[35] S. Wirojanupatump, P.H. Shipway, Adirect comparison of wet and dry abrasion
behaviour of mild steel, Wear 233/235 (1999) 655665.
[36] E. Medvedovski, Wear-resistant engineering ceramics, Wear 249 (2001)
821828.
[37] M. Vite, M. Castillo, L.H. Hernandez, G. Villa, I.H. Cruz, D. Stephane, Dry and wet
abrasive resistance of Inconel 600 and stellite, Wear 258 (2005) 7076.
[38] D.B. Marshall, B.R. Lawn (Eds.), Indentation of Brittle Materials. Micro-
indentation techniques in Materials Science, ASTM, 1986.

S-ar putea să vă placă și