You are on page 1of 20

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/325829340

Effect of Abrasive Particle Size on Friction and Wear Behaviour of HVOF


Sprayed WC-10Co-4Cr coating

Article  in  Materials Research Express · June 2018


DOI: 10.1088/2053-1591/aacd39

CITATIONS READS

3 191

3 authors:

Bikramjyoti Sahariah Nitesh Vashishtha


Indian Institute of Technology Guwahati Metallizing Equipment Co. Pvt. Ltd. Jodhpur India
2 PUBLICATIONS   5 CITATIONS    19 PUBLICATIONS   126 CITATIONS   

SEE PROFILE SEE PROFILE

Sanjay Sapate
Visvesvaraya National Institute of Technology
50 PUBLICATIONS   520 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Effect of solid lubricant addition on friction and wear of plasma sprayed ceramic coatings View project

Modified chromium carbide based coatings for wear resistant applications View project

All content following this page was uploaded by Nitesh Vashishtha on 22 September 2018.

The user has requested enhancement of the downloaded file.


Materials Research Express

PAPER Related content


- Friction and wear behaviour of plasma
Effect of abrasive particle size on friction and wear sprayed Cr2O3-TiO2 coating
Pranay Bagde, S G Sapate, R K Khatirkar
behaviour of HVOF sprayed WC-10Co-4Cr coating et al.

- Mechanical property changes in HVOF


sprayed nano-structured WC-
To cite this article: Bikram Jyoti Sahariah et al 2018 Mater. Res. Express 5 066424 17wt.%Ni(80/20)Cr coating with varying
substrate roughness
Tarek A Ben Mahmud, Gobinda C Saha
and Tahir I Khan

- Tribo-corrosion of coatings: a review


View the article online for updates and enhancements. Robert J K Wood

This content was downloaded from IP address 89.202.245.164 on 02/07/2018 at 09:41


Mater. Res. Express 5 (2018) 066424 https://doi.org/10.1088/2053-1591/aacd39

PAPER

Effect of abrasive particle size on friction and wear behaviour of HVOF


RECEIVED
22 May 2018
sprayed WC-10Co-4Cr coating
REVISED
7 June 2018
ACCEPTED FOR PUBLICATION
Bikram Jyoti Sahariah, Nitesh Vashishtha1 and S G Sapate
18 June 2018 Department of Metallurgical and Materials Engineering, Visvesvaraya National Institute of Technology (VNIT) South Ambazari Road,
PUBLISHED Nagpur- 440010, India
1
27 June 2018 Authors to whom any correspondence should be addressed.
E-mail: niteshsharma282@gmail.com

Keywords: friction, coating, abrasion, particle size, fracture, tribo-oxidation

Abstract
The objective of the present work is to investigate the effect of abrasive particle size on friction and
abrasive wear performance of HVOF sprayed WC-10Co-4Cr coating. WC-10Co-4Cr coating was
deposited by HVOF technique on 316 SS substrate. Friction coefficient and abrasive wear rates were
measured with varying abrasive particle size, load and sliding speed. The microscopic observations,
XRD and XPS analyses of worn surface were used to analyze friction and wear data. The increase in
abrasive wear rate with an increase in load and abrasive particle size was attributed to the transition in
wear mechanisms. The variation in friction coefficient with abrasive particle size as a function of load
and sliding speed was due to change in the contribution of adhesion and abrasion, and varying degree
of tribo oxidation. With increasing sliding speed abrasive wear rate and friction coefficient decreased,
which was due to the dominance of oxidative wear. The prevalence of fracture dominated mechanical
wear processes led to an increased coefficient of friction at higher load.

1. Introduction

Thermal spray coatings are used for prevention against wear, corrosion and high temperature. Amongst the
different thermal spraying techniques, HVOF (High Velocity Oxy-Fuel) has a lower flame temperature and
higher particle velocity, which produces coatings with high density and lower degree of phase transformation,
which can occur during the spraying process [1]. HVOF sprayed tungsten carbide based coatings have been
widely used for improving life of industrial components subjected to sliding wear, abrasion and erosion in
aviation, automotive, pulp and paper, cement and power generation industries.
Typical applications of WC-Co-Cr coating involving abrasive wear includes aircraft landing gears, drill bit
components, rotors in powder mixture and fan blades in cement industries carrying particulate matters [2–6].
Due to addition of chromium, WC-Co-Cr coating offers superior cohesive strength, toughness as compared to
conventional WC-Co coating, which makes them suitable for applications involving severe wear conditions at
high temperatures. WC-Co-Cr coating also exhibits relatively less decarburization of WC, during spraying
[7–10]. The decarburization of WC had an adverse effect on wear resistance [11, 12]. In HVOF sprayed WC-Co-
Cr coating, CoCr serves as a binder to WC particles.
As reported by Murthy et al [13] WC-CoCr coating has better matrix properties and high hardness of WC
particles, which leads to better abrasive wear resistant compared to other coatings. The CoCr matrix has
relatively higher ductility and wettability towards WC particles. The presence of Cr reduces metallic tungsten
formation and decomposition of WC [10, 14]. The grain boundaries of WC are stabilized and WC skeleton is
strengthened due to the Cr addition, as it delays the shrinkage and thus making deformation of WC skeleton
difficult. The temperature at which liquid formation takes place is also lowered by the presence of Cr [15].
The structure and mechanical properties of WC-Co and WC-Co-Cr coatings like hardness, elastic modulus
and fracture toughness are influenced by spraying method, process variables, feedstock size and volume fraction
of carbides and the properties of binder [16–22]. The decomposition of WC forms W2C and η-phase (CoxWyC)

© 2018 IOP Publishing Ltd


Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

which are brittle in nature and deteriorate the coating properties [23]. The wear resistance and fracture
toughness of the coating are reduced due to the presence of brittle phases (W2C, WxCoyC). The WC particle size
should be optimum of about 1.2 μm as reported by Gong et al [3]. As the WC particle size decreases, the level of
decarburization (ratio of W2C/WC) increases [24]. With the decrease in carbide particle size, an increase in
surface area favours formation of W2C phase along the grain boundary of WC particles. With the increase in
carbide particle size, an increase in binder mean free path leads to easy removal of binder and subsequently an
increase in wear rate due to increased pull out of carbide particles, during wear.
The important variables affecting abrasive wear resistance of thermally sprayed coatings are properties of
abrasive particles such as hardness, size and shape, operating parameters such as sliding speed and the load
[25, 26]. WC-Co coating exhibited highest wear rate with SiC abrasive particles followed by that with alumina
and silica abrasives [27]. With the increase in the angularity of abrasive particle, as the proportion of material
removed by micro cutting increases, as opposed to ploughing with spherical particles, leading to an increased
abrasive wear rate. The abrasive wear rate increases as the abrasive particle size increases, but beyond a critical
abrasive particle size of around 150 μm, the wear rate remains more or less constant, which is referred as particle
size effect [28–30]. Numerous explanations have been offered in the past to explain ‘particle size effect’ in
abrasive wear. The reduction in wear rate with the decrease in abrasive particle size was explained with respect to
fracture and fragmentation of particles, capping and clogging, reduction in efficiency of cutting by abrasive
particles, and with as scale of deformation decreases, nucleation of dislocations become difficult in relatively
smaller volume [25, 28].
Qi et al noted an increase in mass loss of more than six times when abrasive particle size was increased from
30 μm to 200 μm for TiC-Ni based HVOF coating and it was also observed that microcrystalline coating
performed better under benign wear conditions while wear resistance of nano crystalline coating was better
under severe wear conditions [31]. The observations due to Sharma et al indicate that abrasion rate of Co-Cr
coating was dominated by abrasive particle size at higher loads when compared to that at lower loads [4]. The
finer abrasive particles exhibited more fracture and fragmentation whereas shelling and attrition was favoured
for coarser abrasive particles and cutting efficiency of abrasive particles was strongly influenced by the density of
abrasive particles [32]. It was noted that three body abrasion rate of WC-Co-Cr coating, increased by a factor of
nearly three for increase in abrasive particle size 4.5 μm to 180 μm. The abrasive wear rates were rationalized
using severity of contact parameter, to deduce transition in abrasion rates [26]. It was concluded by Usmani et al
that carbide size had no significant influence on coefficient of friction (COF) of WC-Co coating in abrasive wear
[15]. The different wear regimes from mild to medium and severe for carbide based coatings were because of
transition in wear mechanisms from plastic deformation, micro fracture of carbides to delamination cracking,
inter granular fracture and splat fracture [25, 33, 34]. An increase in COF for increase in load was noted whereas
abrasive wear rate and COF rise decreased for rise in sliding velocity. The tribo film of lower shear strength
resulted in lowering of wear rate and COF, although magnitude of decrease was influenced by atmosphere,
nature of tribo film, its adherence with the substrate and wear mode [7, 25, 35–38].
Most of the previous studies were focused on sliding wear of WC-Co-Cr coating and a little work has been
reported to study friction and abrasive wear behavior of WC-Co-Cr coating. Effect of abrasive particle size on
wear and friction behavior of WC-Co-Cr coating is not studied previously. The aim of present study is to
investigate the effect of abrasive particle size on friction and abrasive response of HVOF sprayed WC-Co-Cr
coatings under the conditions of varying sliding speed and load.

2. Experimental

2.1. Coating deposition and mechanical properties


Coating powder having a composition of 86%WC, 10%Co, 4%Cr (WOKA 3652, Sulzer Metco, Germany)
manufactured by sintering and agglomeration route having a particle size of −45 μm to +15 μm (spherical
shape) was deposited on 316 stainless steel substrate by High Velocity Oxy Fuel Spray (HVOF) technique. The
thickness of the coating on the substrate was around 350 μm. The gun used for coating deposition was HV-50,
FST, Netherlands. Before coating deposition the substrate surface was grit blasted using alumina grits (size
600–680 μm) with a pressure of 3 kg cm−2. HVOF spraying parameters employed in coating deposition are
specified in table 1. Microhardness was measured by Vickers Indentation hardness tester at 300 g load. In order
to determine the fracture toughness of coating, indentation fracture toughness (IFT) technique was estimated
using equation [25],
⎛ P ⎞ ⎛ a⎞
Kc = 0.079⎜ 3 2 ⎟ log ⎜4.5 ⎟ (1)
⎝a ⎠ ⎝ c⎠

2
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Table 1. HVOF spray deposition parameters (34).

Powder feed rate (g min−1) 45


Fuel supply (l hr−1), Kerosene 20
Oxygen supply (l min−1) 850
Carrier gas supply(l min−1), Nitrogen 12
Coolant supply (l min−1), Water 31–32
Combustion pressure (Bar) 6–6.5
Standoff distance (mm) 360

Figure 1. Fresh SiC abarsive paper of different grits (a) 36 μm (b) 50 μm (c) 100 μm (d) 150 μm.

Where KC=fracture toughness (MPa m ), P=load (mN), a=half-length of diagonal, l=length of crack
and c=a+l. The limit of applicability of IFT equation is 0.6„c/a<4.5. An average of five values is reported
in results.

2.2. Microstructural characterization


The microstructural studies on the coated samples were carried out by JEOL 6380A scanning electron
microscopy (SEM) and EDS. PANalytical X’Pert Pro MPD diffractometer (CuKα radiation) was used for XRD
analysis. For metallography, the samples were polished with SiC emery papers of grit 600, 800, 1200 and 1500
and then cloth polished with alumina suspension. The polished samples were ultrasonically cleaned and then
etched for 15 s using Murakami’s reagent (K3Fe(CN)6+NaOH+distilled water in 1:1:10 ratio). The
characterization of worn samples, worn papers and wear debris was also carried out using SEM, EDS and XRD.

2.3. Abrasive wear testing and analysis of worn surface


SiC particles having size of 36, 50, 100 and 150 μm were used as abrasive medium. SiC particles had an angular
shape as observed in SEM images in figures 1(a)–(d). The abrasive wear tests were carried using Pin on Disk
Tribometer (make, TR- 20LE, Ducom, India). Specimens having dimensions of 10 mm diameter and 20 mm
length were used for wear testing. The specimen preparation and other details of abrasive wear testing have been
described elsewhere [7]. The test parameters for abrasive wear testing were listed in table 2.
The worn-out samples were ultrasonically cleaned and observed using SEM to study morphology of worn
surfaces. The phase analysis of worn surface and wear debris was carried by XRD XPS analysis was also carried on
worn sample to confirm the composition of tribo-oxide film by XPS, ESCA+(Oxford Instrument, Germany)

3
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Table 2. Two body abrasive wear test


parameters.

Parameter Value

Sliding velocity (m s−1) 0.25, 1.0


Abrasive particle size (μm) 36, 50, 100, 150
Load (N) 30, 60
Sliding distance (m) 1256
Temperature (°C) 25
Relative humidity (%) 25–30

using Al-kα radiation having hυ=1486.7 eV. The binding energy accuracy of 0.1 eV is used for analysis.
Carbon peak of 284.6 eV with binding energy is taken as reference. XPS peaks were fitted using software XPS
peak 4.1 and the binding energy of phases was analyzed by using NIST standard database 20 (version 4.1).

3. Results

3.1. Microstructure and mechanical properties


Sintered and agglomerated WC-10 Co-4 Cr powder having spherical shape consisting of CoCr binder and WC
grains can been seen in figure 2(a). Figure 2(b) shows lamellar microstructure of WC-10 Co- 4Cr coating. EDS
spectra as displayed in figures 2(c) and (d) revealed Co, Cr W and C elements in powder as well as in deposited
coating. Uniform distribution of WC grains in CoCr binder matrix was forming a dense coating was also
observed. Bright regions were identified as WC particles surrounding metallic CoCr binder phase (dark region).
Good mechanical interlocking and adhesion with substrate is aided by higher kinetic energy of impinging
particles which helped in flattening of semi -molten/molten particle on the substrate. The decarburization of
WC to W2C depends upon process parameters. The brighter rim around WC particles can be identified as W2C
particles [11, 19, 39]. No micro cracking was seen on the coating surface. XRD spectra of the coating can be seen
in figure 2(e). The major peak of WC along with peaks CoCr were detected in XRD analysis of deposited
coatings. While a peak of W2C was also detected which can be attributed to decarburization reaction during
spraying process, whereas Co peak was not detected possibly due to formation of amorphous or nano crystalline
phase formation [3], as supported by a broad hump observed at 2θ angle of 37°–45° [40]. In addition, a ternary
or eta phase i.e. CoxWyCz was also detected, which forms due to dissolution of W2C and W in the binder phase
[11, 40]. WC-10 Co-4 Cr coating formed good mechanical interlocking with substrate as seen in figure 2(f).
SEM micrographs from polished cross-section were analyzed using image-J software for calculation of
porosity, mean free path and carbide size. Mean free path of CoCr binder was 0.32 μm and WC particle size was
1.5±0.5 μm whereas coating porosity was 1.1%. The microhardness of coating was 1148±100 HV and the
indentation fracture toughness was 5.1±0.7 MPa√m whereas average surface roughness of coating surface
was 5.28 μm. [7]. SEM micrograph of indentation used for calculation of fracture toughness is shown in
figure 2(g). Indentation diagonal and crack length was measured using image analysis software (image-J) on
SEM micrographs of indentation zone.

3.2. Friction coefficient and wear rate


COF versus sliding distance graphs are shown in figures 3(a)–(d). Average COF values at a sliding speed of
0.25 m s−1 and 30 N for abrasive particle sizes of 36, 50, 100, 150 μm were 0.28, 0.332, 0.284, and 0.334,
respectively. The respective COF at 1 ms−1 were 0.216, 0.271, 0.232 and 0.274. At a load of 60 N and sliding
speed of 0.25 m s−1 COF values with increase in abrasive particle size were 0.333, 0.4, 0.37, 0.403 whereas
respective COF at sliding speed of 1 ms−1 were 0.263, 0.34, 0.307, and 0.353. It can be observed from figure 3 that
COF increased in the beginning and then decreased and after that it attains a steady state regime at around
400–500 m of sliding distance. In the beginning, the contact between the coating surface and the abrasive
particles takes place, at a very few of asperity junctions, due to which the contact stress increases leading to
increase in COF. As the sliding distance increases, due to deformation at asperity junctions the real contact area
increases, as a result, it leads to decrease in contact stress and COF. It was also observed that COF then increases
and become stabilized at around 500 m of sliding distance. The increase and stabilization of COF for increasing
sliding distance can be explained in terms of relative contribution of adhesion and abrasion component of
friction, formation and delamination of tribo layer with variation in load and sliding speed, which is discussed in
section 4.2.3.
The effect of abrasive particle size, load and sliding speed on coefficient of friction can be observed from
figures 4(a) and (b). As can be seen, COF increased for increase in load from 30 to 60 N. At 30 N and 60 N loads,

4
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 2. SEM micrographs (a) Coating powder (b) Microstructure of sprayed coating (c) EDS analysis of powder (d) EDS analysis of
sprayed coating surface (e) XRD spectra of sprayed coating surface (f) SEM micrograph of coating cross section (g) Vickers indentation
SEM micrograph.

with an increase in abrasive particle size from 36 to 50 μm, COF increased followed by a decrease to abrasive
particle size of 100 μm and then COF rises for increase in abrasive particle size to 150 μm. Irrespective of the
load, relatively lower COF values were observed for rise in sliding velocity (0.25 to 1 m s−1), as seen in

5
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 3. COF versus sliding distance plot at 0.25 m s−1 (a) 30 N and (b) 60 N at 1 m s−1 (c) 30 N and (d) 60 N.

Figure 4. Coefficient of friction versus particle size at (a) 0.25 m s−1 (b) 1 m s−1.

figures 4(a), (b). At 0.25 m s−1 sliding speed, the increase in COF was in the range of 19%–30% when the load
was increases from 30 to 60 N within the abrasive particle size range of 36 to 150 μm, whereas at a sliding speed of
1 m s−1, the respective value was in the range of 22%–32%. Similarly, the decrease in COF with for an increase in
sliding speed from 0.25 to 1 m s−1, was in the range of 18%–23% at 30 N load whereas at 60 N load, the
reduction in COF was in the range of 12%–21%, for abrasive particle size range of 36–150 μm.
Effect of abrasive particle size (36, 50, 100, 150 μm) on the wear rate of WC-10 Co- 4Cr coating for various
sliding velocities and loads is displayed in figures 5(a), (b). The abrasive wear rates at 0.25 m s−1 sliding speed and
30 N load were 2.3, 3.5, 6.9 and 7.5×10−4 mm3 m−1 for abrasive particle size of 36, 50, 100 and 150 μm,
respectively. The respective values of at 1 ms−1 sliding speed were observed to be 1.8, 2.2, 4 and 4.4×10−4 mm3
m−1. At 60 N load, wear rates were 3.5, 5.5, 10.8, 12×10−4 mm3 m−1 and 2.4, 4.2, 8.2 and

6
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 5. Abrasive wear rate versus particle size at (a) 0.25 m s−1 (b) 1 m s−1.

8.8×10−4 mm3 m−1, for 0.25 m s−1 and 1 m s−1 sliding speeds, respectively. Thus, with increase in abrasive
particle size, abrasive wear rates increased by a factor of 2.5–4 depending on the load and the sliding speed. The
abrasive wear rate increased till abrasive particle size of 100 μm and then exhibited a relatively small increase to
150 μm abrasive particle size as can be seen from figure 5.
The increase in abrasive wear rates at 60 N were around 1.5–1.6 times as compared to those observed at 30 N
at a sliding speed of 0.25 m s−1, whereas at a sliding speed of 1 m s−1, the respective increase was 1.4–2 times,
depending upon the abrasive particle size (36–150 μm). Irrespective of the load, abrasive wear rates decreased
for rise in sliding speeds from 0.25 to 1 m s−1. At a sliding speed of 0.25 m s−1, the decrease in abrasive wear rates
was in the range of 21%–41% whereas for sliding speed of 1 ms−1, the decrease was in the range of 23%–32%,
depending on the abrasive particle size.

4. Discussion

4.1. Wear behavior of WC- 10Co-4 Cr coating


The variation in abrasive wear rates with an increase in abrasive particle size, load and sliding speed can be
explained on the basis of material removal mechanism and tribo oxidation effects. SEM images of worn samples
at different test conditions showed significant differences in the morphology of abraded surfaces, as shown in
figure 6. At a load of 30 N and sliding speed of 0.25 m s−1 for abrasive particle size less than 100 μm, the worn
surfaces were relatively smooth. Due to shear and normal stresses, small displacement of carbide grains was
observed along with micro fracture of carbide grains, (figure 6(a)) assisted by hertzian elastic contact. The
removal of binder phase occur by repeated plastic deformation and micro grooving mechanism [25, 34], as
shown in figure 6(b). With finer abrasive particles and at lower load of 30 N, WC-10Co- 4Cr coating displayed
relatively lower wear rates, as shown in figure 5. At 60 N load, an increase in the abrasive wear rate was due to
increased depth of grooves caused by SiC particles in the soft binder phase, extrusion and removal of ductile
binder phase, cracking and fracture of carbide grains [41]. Similarly, assisted by easy removal of binder phase at
higher load, the unsupported carbide particles can be removed either by their bodily removal or by pull out, as
can be observed from figures 6(c) and (d). When the contact stress at the asperity junction exceeds the yield limit
of the material, micro cracking and fracture was favoured by accumulation of twins and dislocations [42]. With
coarse abrasive particle size (100–150 μm) and for increasing load, wear mechanisms displayed a shift from
plastic deformation to mechanisms dominated by fracture. The increased abrasive wear rates at higher load and
coarser abrasive particles were attributed to material removal mechanisms predominated by delamination
cracking, intergranular cracking, nucleation of cracks at inter splat boundaries and their interlinkage, leading to
inter splat and subsurface splat fracture [25, 34], as observed in figures 6(e)–(h).
The reduction in the abrasive wear rate with decreasing abrasive particle size can be explained by ‘particle
size effect’. With finer abrasive particles, the decrease in abrasive wear rate can be attributed to capping, clogging,
fracture and fragmentation of abrasive particles [28, 30, 32]. The tip of abrasive particle gets covered with wear
debris (capping), which reduces its cutting efficiency, accumulation of debris particles between the spaces of
abrasive particles (clogging), leading to reduction in the load carried by each abrasive particle. Similarly, fracture
and fragmentation of particle leads to reduction in attack angle posed by abrasive particles causing decrease in
material removal rate [43]. In case of two phase materials, like cermets, both abrasive wear mechanisms; i.e.

7
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 6. SEM of wornout sample at 36 μm (a) 30 N, 0.25 m s−1 (b) 30 N, 1 m s−1, (c) 60 N, 0.25 m s−1 at 50 μm (d) 60 N 1 m s−1 , at
100 μm (e) 60 N, 1 m s−1 (f) 30 N 1 m s−1 , and at 150 μm (g) 60 N, 0.25 m s−1 (h) 60 N, 1 m s−1.

plastic deformation and fracture contribute to total abrasive wear rate. The contribution of each mechanism is
affected by the abrasive particle size and load. With finer abrasive particles and lower loads, predominant
contribution arises due to plastic deformation of the binder by ploughing and micro cutting and fatigue assisted
micro fracture of carbide grains. With coarser abrasive particles and at higher loads, material removal rate was
determined by wear mechanisms dominated by fracture, as shown in SEM micrographs of worn surfaces
(figures 6(g), (h)), which could support the above observations. The other reason is that the tribo film formed on
the worn surface cannot be penetrated easily by relatively finer particles whereas with coarser particles, tribo film
is removed by delamination cracking, which results in increased abrasive wear rate with increasing abrasive
particle size. The tribo film formation and its effect on abrasive wear behaviour of WC- 10Co- 4Cr coating is
explained as under.
The abrasive wear rate of WC-10 Co- 4 Cr coating decreased at higher sliding speed, which could be ascribed
to phenomena of tribo oxidation [7]. As the sliding velocity is increased (0.25–1 m s−1) besides mechanical wear

8
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 7. EDS analysis of selected regions from SEM images of worn coating surfaces at different test conditions.

processes, oxidative wear mechanisms play crucial role in modifying wear and friction behaviour of the coating.
As wear proceeds, plastic deformation at asperity junction generates wear debris particle. With further wear
cycle, the fragmentation and fracture of wear debris particles results in very fine particles; some these wear debris
particles are swept away while frictional heating leads to oxidation of debris particles in the wear tracks. The
oxidation, agglomeration and subsequent compaction of fine debris particles forms tribo film (tribo oxide film)
on abraded surface [25, 44]. The tribo film (as seen in figures 6(e)–(h)), with low shear strength acts like solid
lubricant and preventing the contact at the asperity junctions and reduces not only the abrasive wear rate but also
the friction coefficient [7]. The magnitude of decrease in wear rate for increase in sliding speed (0.25 to 1 m s−1),
increased with the increase in abrasive particles size from 36 μm (22%) to 150 μm (41%) at 30 N load, whereas at
60 N, the respective reduction in wear rate was 31% and 26% with 36 μm and 150 μm abrasive particle size,
respectively. The variation of friction and abrasive wear response with abrasive particle size, load and sliding

9
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 8. Plot between flash temperature and abrasive particle size.

speed is explained by the effect of abrasive particle size on tribo oxidation, which is discussed in next section.
EDS analysis of selected regions from SEM images of worn coating surfaces at different test conditions is given in
figure 7.

4.2. Abrasive particle size and tribo oxidation


4.2.1. Flash temperature
In order to evaluate the effect of abrasive particle size on tribo oxidation, flash temperatures were estimated
under varying abrasive wear conditions. An increase in temperature (flash temperature) as a result of frictional
heating during wear depends on sliding speed, load and COF and the real area of contact. The flash temperature
can be estimated by the following expression [25, 45, 46],
m PV
DT = (2)
4J (Ka + Kcoat) a
⎛ P ⎞½
Where, a = ⎜ ⎟ (3)
⎝ p Hcoat ⎠

where, ΔT=rise in flash temperature (°C), P=applied load (N), μ=coefficient of friction, V=speed (m
s−1), J=Joule’s constant (1), Kcoat and Ka are thermal conductivity of coating and SiC, a=real area of contact
(m), Hcoat=coating hardness (N m−2).
Figure 8 shows graph of estimated flash temperatures vs. abrasive particle size. The flash temperatures at
30 N load and at a sliding speed of 0.25 m s−1 and at abrasive particle sizes of 36, 50, 100, 150 μm were 256, 304,
260 and 306 °C, respectively. At 30 N load and at 1 m s−1 speed, the flash temperatures were 792, 994, 851 and
1005 °C. At a load of 60 N, flash temperatures were 431, 518, 479 and 522 °C and 1364, 1764, 1592, 1831 °C, at
sliding speed of 0.25 m s−1 and 1 m s−1. An increase in abrasive particle size, load and sliding speed resulted in
rise in flash temperature from as low as 256 °C to as high as 1831 °C was , indicating that severity of tribo
oxidation increased as abrasive wear condition became more severe.
Apart from microscopic observation of worn surfaces, the tribo oxidation of the worn surface was also
indicated by EDS analyses of worn surfaces as seen in figure 7. As shown in table 3, an increase in oxygen
concentration with increased abrasive particle size, sliding speed and load was indicative of an increase in the
extent of tribo oxidation. The coverage of worn surface by tribo-oxide film increased with rise in load, sliding
speed and abrasive particle size as revealed by SEM observations of abraded surfaces, in figures 6(a)–(h). The
coverage of worn out surface by tribo-film was crucial in decreasing abrasive wear rate, as pointed out previously
[25]. The effectiveness of tribo film in altering wear and friction behaviour is influenced by its adherence with the
substrate, composition of tribo-film, its hardness, geometry of contacting surfaces and wear mode, load and
sliding speed [25, 35]. At higher load, sliding speed and with coarser abrasive particles, medium to severe
oxidation of worn surface was observed, which was supported by estimated flash temperatures, shown in
figures 6(e) and (h). At lower load and with finer abrasive particles, mild oxidation prevails and tribo film is
relatively porous and is removed by plastic shearing. The process of tribo oxidation is dynamic one, involving
formation, removal and reformation of tribo film, referred as ‘oxidation- scrape-reoxidation’ [28]. At relatively

10
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Table 3. EDS analysis of wornout surfaces at different load, speed and particle size.

Sliding speed (m s−1) Load (N) % Composition

0.25 30 N Particle size (μm) W C Co Cr O


36 72.08 6.08 10.4 4.41 7.02
50 67.33 6.69 8.86 5.87 11.88
100 59.64 9.05 9.09 7.52 14.69
150 58.01 9.54 8.57 7.49 16.36
60 N 36 70.23 5.11 9.3 5.96 9.37
50 65.36 7.57 7.12 5.64 14.29
100 58.03 8.99 7.93 6.97 17.54
150 57.11 8.88 6.27 6.42 21.3

1.0 30 N 36 68.09 10.74 5.85 3.43 11.89


50 57.57 8.77 9.71 4.3 19.65
100 58.43 6.76 6.39 5.92 22.38
150 51.16 7.33 8.51 4.01 28.97
60 N 36 61.57 7.48 10.23 6.4 14.32
50 59.3 4.99 8.08 3.76 23.87
100 52.32 7.77 5.19 3.89 30.8
150 47.98 6.67 5.14 3.76 36.44

higher load and for coarser abrasive particles, tribo oxide film is broken by indentation caused by SiC particles
and delamination cracking, as a result of which tribo film is not able to provide the protection to worn surfaces
with corresponding rise in COF (figure 6(h)). The above explanation could also justify the different trends in
magnitude of variation in abrasive wear rate when the abrasive particle size and load was increased.

4.2.2. Characterization of worn surface


To ascertain the composition of tribofilm, XRD analyses of worn surfaces and wear debris was carried out. XRD
analyses of abraded surfaces are shown in figure 9. At lower abrasive particle size (36 and 50 μm) and at lower
speed (0.25 ms−1), XRD spectra revealed peaks of oxides of W (WO3), Co (CoO, Co3O4), CoWO4 along with
WC and W2C as observed from figures 9(a) and (b). For abrasive size of 100 μm and at same speed, additional
peak of spinel phase, CoCr2O4 was also detected, as seen in figure 9(c). At sliding speed of 1 ms−1, XRD peaks
revealed oxides of Cr (CrO2, CrO3) and W (WO2, W3O8) apart from phases detected at lower sliding speed as
seen in figures 9(d)–(f). The phase analysis of tribo film as revealed by XRD analysis was indicative of an increase
in the extent of tribo oxidation with increased load, sliding speed and abrasive particle size, which has resulted in
altering wear and friction properties of WC-10Co-4 Cr coating. Additionally, worn surface at 60 N load with
50 μm abrasive particle size and velocity of 0.25 ms−1 was analyzed by XPS, which revealed strong peaks of
oxides of W, Cr, Co and binder phase of Co and Cr, as shown in figures 10(a)–(e). The respective binding
energies of various oxides and hydroxides is indicated, as per the following. Tribo-oxide film of WO3 (36,
38.1 eV) has been revealed by high- spectral peak of W4f. A peak of Co2p revealed the presence of CoO (781.2
and 796.5 eV) and Co(OH)2 (803.7 eV). The peak of Cr2p revealed the presence of tribo-oxide film of Cr2O3
(576.1 eV), CrO3 (578.3 and 580.1 eV) and spinel phase, CoCr2O4 (586.1 eV). The peak of Si2p indicated the
formation of SiO2.
The formation of surface films of tribo oxides of W, Co and Cr can be given by chemical reactions [47–53] as
given in table 4. The standard Gibbs free energy of formation is indicative of relative stability of various phases,
more negative is the free energy, and greater is the stability of oxides. The formation of Co(OH)2 and SiO2 could
be attributed to tribo chemical reaction of Co and fragmented SiC particles with environment, i.e. humidity
(H2O). XPS spectra also revealed peals of binder phase i.e. Co and Cr. Due to being ductile and soft cobalt helps
in densification of tribo film by acting as a binder, thus improving its adhesion with the worn surface [44]. EDS
analysis of worn surface (figure 7 and table 3) also indicated enrichment of cobalt in the tribo film on the worn
surface. The formation of pure cobalt oxide is favoured when the cobalt is available in the abundant quantity and
when the flash temperature rise temperature is above 600 °C–700 °C, which justifies the present of cobalt partly
in metallic form and remaining cobalt forms oxides, i.e. CoO and Co3O4 [54].
Thus, cobalt oxide peaks (CoO and Co3O4) can be ascribed to oxidation of extruded cobalt and its further
oxidation to Co3O4. Co3O4 is thermodynamically more stable and is more effective in decrease of friction
coefficient than CoO, which is metastable oxide and tungsten oxides [51, 53]. The oxidation of WC results in
formation of WO3 and WO2. Magneli phase oxides (WOx) facilitates easy shear of crystallographic planes and
helps in reduction of COF [55, 56]. Cr2O3 is more stable but its relatively hard and brittle as compared to other

11
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 9. XRD analysis of wornout samples at 0.25 m s−1 (a) 30 N 50 μm (b) 60 N 36 μm (c) 60 N 100 μm, at 1 m s−1 (d) 30 N
100 μm (e) 60 N 50 μm (f) 60 N 150 μm.

oxides and is not considered to be effective because of it lower adhesion with the coating surface, as pointed by
Clemow [57].

4.2.3. Analysis of wear debris and worn out abrasive paper


Figure 11 shows SEM micrographs of wear debris particles, which reveal difference in the morphology of wear
debris particles with change in load and abrasive particle size. It also showed fragmented and fractured SiC
particles, WC particles, oxides of W, Co, Cr and Si, as confirmed by XRD analysis which is shown in figure 12. At
60 N load and sliding speed of 1 m s−1 with abrasive particle size of 36 μm, XRD spectra revealed peaks of WC,
SiC, SiO2, oxides of W (WO2, WO3), oxide of Co (CoO) with CoWO4 and spinel phase CoCr2O4 (figure 12(a)).
With abrasive size of 50 μm, along with the above phases, peaks of Co3O4 and W3O8 were also detected as shown
in figure 12(b). With 100 μm abrasive particle size, additional peak of CrO3 was also seen (figure 12(c)), whereas
with 150 μm abrasive particle size, along with above mentioned phases, CoWO4 peak was also detected as shown
in figure 12(d). SiC peaks confirmed fracture and fragmentation of SiC abrasive particle during abrasive wear
test. SEM images of the SiC abrasive paper with different particle size of 36, 50, 100 and 150 μm and at 60 N load
and speed of 0.25 m s−1 are displayed in figures 13(a)–(d). EDS spectra of wornout papers at higher load can be
seen from figures 13(e)–(f). SEM images reveal capping, clogging, attrition, fragmentation and fracture and

12
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 10. XPS analysis of worn coating surface at 0.25 m s−1 60 N for 50 μm abrasive particle size.

Table 4. Tribo-oxidation reactions [47–53].

No. Equation ΔG298K (KJ mol−1)

1 WC + 5 2O2  WO3 + CO2 −764.12


2 WC + O2  WO2 + C −533.87
3 2Co + O2  2CoO −214.22
4 3CoO + 1/2O2  Co3 O4 −794.96
5 Cr + 3 2O2  CrO3 −273.46
6 2Cr + 3 2O2  Cr2 O3 −1053.11
7 SiC + 2H2 O  SiO2 + CH4 −372.9
8 Co + H2 O  Co (OH )2 −454.38

removal of abrasive particles (shelling). The clogging, attrition and edge rounding of abrasives, was more
prevalent in case of relatively fine abrasive particles and at lower load while at higher load (60 N), fracture,
fragmentation and removal of abrasive particles (shelling) was observed.
The friction response of WC- 10Co-4 Cr coating can be explained in terms of relative contribution of
adhesion and abrasion component of friction and oxidative and mechanical oxidative wear with change in
abrasive size, load and sliding speed. COF comprises of two components namely; adhesion component and

13
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 11. SEM microgpahs of wear debris at 1 m s−1 sliding speed; at 50 μm particle (a) 30 N (b) 60 N and at 150 μm particle (c) 30 N
(d) 60 N.

Figure 12. XRD analysis of wear debris at 60 N and 1 m s−1 (a) 36 μm (b) 50 μm (c) 100 μm (d) 150 μm.

14
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

Figure 13. SEM micrographs of worn SiC abrasive paper at 1 m s−1 60 N (a) 36 μm (b) 50 μm (c) 100 μm (d) 150 μm; (e), (f) EDS
analysis of selected points.

ploughing component. The adhesion component is dependent on real area of contact while ploughing
component is governed by attack angle of the asperities. The finer SiC abrasive particles exhibited greater
tendency of fracture and fragmentation whereas coarser abrasive particles are vulnerable to attrition and shelling
[32]. With decrease in particle size, the space between abrasive particles decreases and wear debris tend to
accumulate in relatively smaller volume. EDS analysis of worn paper as shown in figures 13(e), (f), indicate
oxygen peaks which confirms accumulation of oxide debris on abrasive paper. With finer abrasive size (below
100 μm) phenomena of clogging dominates in which accumulation and compaction of wear debris occur on
abrasive paper, resulting in enhancement of the adhesion component with corresponding increase in COF as the
load is increased [38, 58]. The friction coefficient exhibited a decrease at 100 μm abrasive particle size, which
could be possibly due to balancing of adhesion and abrasion component and an increase in the degree of tribo
oxidation with increased in abrasive particle size, as indicated by microscopic observations and analysis of
abraded surfaces, as shown in figures 7 and 9. The tribo film acts as solid lubricant and not only separates the
contact at the asperity junction and decreases real area of contact, thus reducing COF, as discussed in section 4.1.
It is interesting to note that the decrease in COF from 50 to 100 μm abrasive particle size was not greater than
15%, which could be attributed to: (a) an increase in the extent of tribo oxidation with increasing abrasive
particle size; (b) equilibrium between formation and removal of tribo oxide film and (c) balancing of adhesion
and abrasion component. It was observed that a further increase in COF when the abrasive particle size is
increased from 100 to 150 μm, was due to combined effect of (a) increased fracture assisted mechanical wear,
increased rate of removal of tribo oxide film by delamination cracking, indentation induced fracture and

15
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

spalling; and (b) reduction in cutting (abrading) efficiency of coarser abrasive particles due to their fracture,
fragmentation and shelling. Previous observations indicate a minimum in COF at abrasive particle size of 35 μm
followed by an increase in COF beyond 35 μm [43]. Due to different relative hardness of the surface (hardness of
abrasive to hardness of surface; Ha/Hs) and the test parameters, it is difficult to make comparison of the findings
of present study with published however, the general trends obtained in change of abrasive wear rates and COF
with abrasive particle size are consistent with previous observations [31, 43, 58].

5. Conclusions

The wear and friction response of HVOF sprayed WC- 10Co-4 Cr coating with varying sliding speed, load and
abrasive particle size was studied in the present investigation. The abrasive particle size effect on wear rate and
COF was rationalized on the basis of particle size effect, tribo oxidation and transition in wear mechanisms,
which was supported by microscopic observations and analysis of worn surface and wear debris. The main
conclusions of the present work are as follows. The increase in abrasive wear rate of WC-10 Co-4Cr coating with
an increase in abrasive particle size and load was due to transition in wear mechanisms from fatigue, plastic
deformation to mechanisms dominated by fracture. The variation of friction coefficient with abrasive particle
size, load and sliding speed was rationalized in terms of change in contribution of adhesion and abrasion
component of COF and varying degree of tribo oxidation. The reduction in abrasive wear rates and COF with
rise in sliding velocity was attributed to lubricious tribo film of oxides W, Co and Cr. The composition, nature
and stability of tribo films altered wear and friction response of coating.

Acknowledgments

The authors are grateful to the Director, VNIT for providing necessary facilities in carrying out this investigation
and TEQIP II grant, FNO.16-6/2015-TS-VII, GOI, MHRD India for financial assistance. The authors are also
thankful to Sprint Testing Solutions laboratory, Mumbai for assistance in XPS measurements.

ORCID iDs

Nitesh Vashishtha https://orcid.org/0000-0003-4620-6699


S G Sapate https://orcid.org/0000-0002-5712-0412

References
[1] Pawlowski L 2007 The Science and Engineering of Thermal Spray Coatings 2nd edn (New York: Wiley)
[2] Bergmann C P and Vicenzi J 2011 Protection Against Erosive Wear Using Thermal Sprayed Cermet (Berlin: Springer)
[3] Gong T, Yao P, Zuo X, Zhang Z, Xiao Y and Zhao L 2016 Influence of WC carbide particle size on the microstructure and abrasive wear
behavior of WC–10Co–4Cr coatings for aircraft landing gear Wear 362–363 135–45
[4] Sharma S 2014 Parametric study of abrasive wear of Co–CrC based flame sprayed coatings by response surface methodology Tribol. Int.
75 39–50
[5] Sapate S G and Roy M 2015 Solid particle erosion of thermal sprayed coatings Thermal Sprayed Coatings and Their Tribological
Performances ed M Roy and J P Davim (Hershey, PA: IGI Global Publications) pp 193–226
[6] Kembaiyan K T and Keshavan K 1995 Combating severe fluid erosion and corrosion of drill bits using thermal spray coatings Wear
186–187 487–92
[7] Vashishtha N, Sapate S G, Gahlot J S and Bagde P 2018 Effect of tribo oxidation on friction and wear behaviour of HVOF sprayed WC–
10Co–4Cr coating Tribol. Lett. 66 56
[8] Basak A K, Achanta S, Bonte M D, Celis J P, Vardavoulias M and Matteazzi P 2007 Effect of Al and Cr addition on tribological behaviour
of HVOF and APS nanostructured WC–Co coatings Trans. of IMF 85 310–5
[9] Berger L M, Vuoristo P, Mantyla T, Kunert W, Lengauer W and Ettmayer P 1996 Microstructure and properties of WC-Co-Cr coatings
Thermal spray: Practical Solutions for Engineering Problem ed C C Berndt (Ohio: ASM International) pp 96–106
[10] Karimi A, Verdon C and Barbezat G 1993 Microstructure and hydroabrasive wear behaviour of high velocity oxy-fuel thermally
sprayed WCCo(Cr) coatings Surf. Coat. Technol. 57 81–9
[11] Stewart D A, Shipway P H and McCartney D G 2000 Microstructural evolution in thermally sprayed WC–Co coatings: comparison
between nanocomposite and conventional starting powder Acta Mater. 48 1596–604
[12] Vashishtha N, Sapate S G, Bagde P and Rathod A B 2018 Effect of heat treatment on friction and abrasive wear behaviour of WC-12Co
and Cr3C2-25NiCr coatings Tribol. Int. 118 381–99
[13] Murthy J K N and Venkataraman B 2006 Abrasive wear behaviour of WC–CoCr and Cr3C2–20(NiCr) deposited by HVOF and
detonation spray processes Surf. Coat. Technol. 200 2642–52
[14] Bolelli G, Lusvarghi L and Barletta M 2009 HVOF-sprayed WC–CoCr coatings on Al alloy: effect of the coating thickness on the
tribological properties Wear 267 944–53
[15] Usmani S, Sampath S, Houck D L and Lee D 1997 Effect of carbide grain size on the sliding and abrasive wear behavior of thermally
sprayed WC-Co coatings Tribol. Trans. 40 470–8

16
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

[16] Bolelli G, Berger L M, Börner T, Koivuluoto H, Lusvarghi L and Lyphout C 2015 Tribology of HVOF and HVAF sprayed WC–10Co4Cr
hardmetal coatings: a comparative assessment Surf. Coat. Technol. 265 125–44
[17] Santana Y Y, Barbera-Sosa L, Caro J, Puchi-Cabrera E S and Staia M H 2008 Mechanical properties and microstructure of WC–10Co–
4Cr and WC–12Co thermal spray coatings deposited by HVOF Surf. Eng. 24 374–82
[18] Zhao L, Maurer M, Fischer F, Dicks R and Lugscheider E 2004 Influence of spray parameters on the particle in-flight properties and the
properties of HVOF coating of WC-CoCr Wear 257 41–6
[19] Jacobs L, Hyland M M and Bonte M D 1998 Comparative study of WC-cermet coatings sprayed via the HVOF and the HVAF process
J. Thermal Spray Technol. 7 213–8
[20] Lee C W, Han J H, Yoon J, Shin M C and Kwun S I 2010 A study on powder mixing for high fracture toughness and wear resistance of
WC–Co–Cr coatings sprayed by HVOF Surf. Coat. Technol. 204 2223–9
[21] Thakur L and Arora N 2013 Sliding and abrasive wear behaviour of WC-CoCr coatings with different carbide sizes J. Mater. Engg.
Perform. 22 574–83
[22] Chivavibul P, Watanabe M, Kuroda S and Shinoda K 2007 Effects of carbide size and Co content on the microstructure and mechanical
properties of HVOF-sprayed WC-Co coatings Surf. Coat. Technol. 202 509–21
[23] Upadhyaya R, Tailor S, Shrivastava S and Modi S C 2017 High performance thermal-sprayed WC-10Co-4Cr coatings in narrow and
complex areas Surf. Engg. 34 412–21
[24] Ghabchi A, Varis T, Turunen E, Suhonen T, Liu X and Hannula S P 2010 Behavior of HVOF WC-10Co4Cr coatings with different
carbide size in fine and coarse particle abrasion J. Thermal Spray Technol. 19 368–77
[25] Vashishtha N and Sapate S G 2017 Abrasive wear maps for high velocity oxy fuel (HVOF) sprayed WC-12Co and Cr3C2−25NiCr
coatings Tribol. Int. 114 290–305
[26] Thakare M R, Wharton J A, Wood R J K and Menger C 2012 Effect of abrasive particle size and the influence of microstructure on the
wear mechanisms in wear-resistant materials Wear 276–277 16–28
[27] Babu P S, Basu B and Sundararajan G 2010 Abrasive wear behavior of detonation sprayed WC–12Co coatings: influence of
decarburization and abrasive characteristics Wear 268 1387–99
[28] Hutchings I M 1992 Friction and wear of Engineering Materials (Oxford: Butterworth-Heinemann)
[29] Tressia G, Penagos J J and Sinatora A 2017 Effect of abrasive particle size on slurry abrasion resistance of austenitic and martensitic
steels Wear 376–377 63–9
[30] Singh K, Khatirkar R K and Sapate S G 2015 Microstructure evolution and abrasive wear behaviour of D2 steel Wear 328–329 206–16
[31] Qi X, Eigen N, Aust E, Gartner F, Klassen T and Bormann R 2006 Two-body abrasive wear of nano- and microcrystalline TiC–Ni-based
thermal spray coatings Surf. Coat. Technol. 200 5037–47
[32] Narayanaswamy B, Hodgson P and Beladi H 2016 Effect of particle characteristics on the two-body abrasive wear behaviour of a
pearlitic steel Wear 354–355 41–52
[33] Gee M G, Gant A and Roebuck B 2007 Wear mechanisms in abrasion and erosion of WC/Co and related hardmetals Wear 263 137–48
[34] Vashishtha N, Khatirkar R K and Sapate S G 2017 Tribological behaviour of HVOF sprayed WC-12Co, WC-10Co-4Cr and
Cr3C2-25NiCr coatings Tribol. Int. 105 55–68
[35] Nicholls J R and Wellman R G 2012 Oxidative wear ed W Robert Bruce, Handbook of Lubrication and Tribology: Theory and Design 2nd
edn (Boca Raton, FL: CRC Press) pp 1–24
[36] Stott F H, Lin D S and Wood G C 1973 The structure and mechanism of formation of the ‘glaze’ oxide layers produced on nickel-based
alloys during wear at high temperatures Corros. Sci. 13 449–69
[37] Pauschitz A, Roy M and Franek F 2008 Mechanisms of sliding wear of metals and alloys at elevated temperatures Tribol. Int. 41 584–602
[38] Bagde P, Sapate S G, Khatirkar R K and Vashishtha N 2018 Friction and abrasive wear behaviour of Al2O3- 13TiO2 and Al2O3 -13TiO2
-Ni Graphite coatings Tribol. Int. 121 353–72
[39] Guilemany J M, Paco J M D, Nutting J and Miguel J R 1999 Characterization of the W2C phase formed during the high velocity oxygen
fuel spraying of a WC+12 Pct Co powder Metall. Mater. Trans. A 30A 1913–21
[40] Wang D, Zhang B, Jia C, Gao F, Yu Y, Chu K, Zhang M and Zhao X 2017 Influence of carbide grain size and crystal characteristics on the
microstructure and mechanical properties of HVOF-sprayed WC-CoCr coatings Int. J. Refract. Met. Hard Mater. 69 138–52
[41] Basse J L 1985 Binder extrusion in sliding wear of WC-Co alloys Wear 105 247–56
[42] Zhao X, An Y, Hou Z and Chen J 2014 Friction and wear behavior of plasma-sprayed Al2O3-13 wt% TiO2 coatings under the
lubrication of liquid paraffin J. Therm. Spray Technol. 23 666–75
[43] Trevisiol C, Jourani A and Bouvier S 2017 Effect of hardness, microstructure, normal load and abrasive size on friction and on wear
behaviour of 35NCD16 steel Wear 388–389 101–11
[44] Yang Q, Senda T and Ohmori A 2003 Effect of carbide grain size on microstructure and sliding wear behavior of HVOF-sprayed WC-
12% Co coatings Wear 254 23–34
[45] Liu Y, Erdemir A and Meletis E I 1996 A study of the wear mechanism of diamond like carbon films Surf. Coat. Technol. 82 48–56
[46] Bagde P, Sapate S G, Khatirkar R K, Vashishtha N and Tailor S 2018 Friction and wear behaviour of plasma sprayed Cr2O3-TiO2 coating
Mater Research Express 5 026410
[47] Dean J A 1979 Lange’s Handbook of Chemistry 12th edn (New York, New York: McGraw-Hill)
[48] Yamanaka K, Mori M and Chiba A 2015 Surface characterisation of Ni-free Co–Cr–W-based dental alloys exposed to high
temperatures and the effects of adding silicon Corros. Sci. 94 411–9
[49] Singhal S C 1975 Effect of water vapor on the oxidation of hot-pressed silicon nitride and silicon carbide J. Americ. Cerm. Soc. 59 81–2
[50] Voitovich V B, Sverdel V V, Voitovich R F and Golovko E I 1996 Oxidation of WC-Co, WC-Ni and WC-Co-Ni hard metals in the
temperature range 500 °C–800 °C Int. J. Refract. Met. Hard Mater. 14 289–95
[51] Chen L, Yi D, Wang B, Liu H and Wu C 2016 Mechanism of the early stages of oxidation of WC–Co cemented carbides Corros. Sci. 103
75–87
[52] Petitto S C, Marsh E M, Carson G A and Langell M 2008 Cobalt oxide surface chemistry: the interaction of CoO (1 0 0), Co3O4 (1 1 0)
and Co3O4 (1 1 1) with oxygen and water J. Molecular Catalysis A: Chemical 281 49–58
[53] Douglass D L and Armijo J S 1971 The influence of manganese and silicon on oxidation behaviour of Co-20Cr Oxidation of Metals 3
185–202
[54] Engqvist H, Hogberg H, Botton G A, Ederyd S and Axen N 2000 Tribofilm formation on cemented carbides in dry sliding conformal
contact Wear 239 219–28
[55] Magneli 1953 Structures of the ReO3-type with recurrent dislocations of atoms: ‘homologous series’ of molybdenum and tungsten
oxides Acta Crystallogr 6 495–500

17
Mater. Res. Express 5 (2018) 066424 B J Sahariah et al

[56] Wesmann J A R, Kuroda S and Espallargas N 2017 The role of oxide tribofilms on friction and wear of different thermally sprayed WC-
CoCr J. Therm. Spray Technol. 26 492–502
[57] Clemow A J T and Daniell B L 1980 The influence of microstructure on the adhesive wear resistance of a Co-Cr-Mo alloy Wear 61
219–31
[58] Jourani A and Bouvier S 2015 Friction and wear mechanisms of 316L stainless steel in dry sliding contact: effect of abrasive particle size
Tribol. Trans. 58 131–9

18

View publication stats

You might also like