Você está na página 1de 9

Correlation between residual stress and abrasive wear of

WC17Co coatings
O.P. Oladijo
a,b,
, A.M. Venter
b,c
, L.A. Cornish
a,b
a
School of Chemical and Metallurgical Engineering, University of the Witwatersrand, Private Bag 3, WITS, 2050, South Africa
b
DST/NRF Centre of Excellence in Strong Materials, South Africa
c
Research & Development Division, NECSA Limited, Pretoria, South Africa
a b s t r a c t a r t i c l e i n f o
Article history:
Received 2 September 2013
Accepted 26 January 2014
Available online 30 January 2014
Keywords:
Residual stress
HVOF
WCCo coating
X-ray diffraction
Abrasive wear
This investigation had been conducted to determine the inuence of residual stresses on the abrasive wear resis-
tance of HVOF thermal spray WC17 wt.% Co coatings, as well as to derive stress relaxation after cutting by wire
electric discharge machining (EDM). The abrasive wear properties of the coatings were characterised using an
ASTM-G65 three body abrasive wear machine with silica sand as the abrasive. The residual stress was measured
by means of X-ray diffraction techniques, on the coated samples before and after the abrasive wear tests. Com-
pressive residual stresses were observed in the surface layer of the large coated samples. However, stress relax-
ation results after cutting into small sizes were distinctly different. There was strong correlation between residual
stresses in the surface layer and abrasive wear resistance, as well as yield strength of a material.
2014 Elsevier Ltd. All rights reserved.
1. Introduction
The choice of a metal substrate for thermal spraying needs adequate
attention in the industrial world. Previous studies based on global char-
acterisation [1,2] has shown that the choice of substrate material for tri-
bological applications of thermal spray coatings relies upon:
ability of substrate material to support the coating.
plasticity during grit-blasting prior to thermal spraying to attain a
critical surface roughness, to provide mechanical interlock as well
as increased contact area.
higher coefcients of thermal expansion than the coating material to
generate a limited degree of compressive residual stress in the coat-
ing. An important step to knowthe particular substrate for a particu-
lar engineering application is to understand more precisely how the
microstructure of the coating affects the coating properties.
Cermet thermal spray coatings are widely used in industry due to
their characteristics such as resistance to abrasion, erosion, high tem-
perature and corrosive atmospheres [3]. The choice of WC over other
carbides is prefer due to the good adhesive characteristic and wettabil-
ity by most binder metals. Cobalt is the most commonly-used binder if
abrasion resistance is required due to its excellent carbide wetting and
adhesion properties [4].
Recently [2,3], many thermal spraying techniques such as air plasma
spraying (APS), high velocity oxyl fuel (HVOF) spraying and vacuum
plasma spraying (VPS) have been applied to deposit WCCo coatings,
although the coatings properties depend on spraying parameters.
HVOF spraying is one of the best methods for depositing conventional
WCCo cermet powder due to the high velocities and lower tempera-
tures experienced by the powder, which results in less decomposition
of the WC during spraying [5,6]. In addition, it offers an effective and
economic method of enhancing wear resistance without compromising
other attributes of the component. Mantyla et al. [7] have evaluated 12%
and 17% cobalt WC/Co coatings deposited by VPS, APS, HVOF and DG
(detonation gun), results had shown that HVOF and DG gave the best
properties and abrasive wear resistance [8]. A microstructural and ana-
lytical study of thermally sprayed WCCo coatings deposited unto aus-
tenitic stainless steel substrate was done [9]. The results showed that
the microstructure of the coating cross section comprised islands, elon-
gated in directions parallel to the substrate. Despite these improve-
ments, microstructure defects within the coating sometimes lower the
adhesive and cohesive strength leading to spalling and fracture [9].
A major factor inuencing the fracture and spalling in thermal spray
coating is the residual stress prole between the coating and the sub-
strate material [10]. The residual stress effects may be either benecial
or detrimental, depending upon the sign and distribution of stresses
with respect to some external factors (e.g. load). Montay et al. [11] iden-
tied that the residual stress gradient in the depth of the top coat, bond
coat and substrate came from three sources. The rst is due to the
thermal and kinetic energy of individual sprayed particles that are inu-
enced by the process itself. The second comes from the thermal expan-
sion mismatch between the coating and the substrate. The third is the
substrate preparation, which inuences the residual stress gradient.
Venter et al. [12] investigated the residual strain associated with WC
Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
Corresponding author at: School of Chemical & Metallurgical Engineering, University
of the Witwatersrand, Private Bag 3, WITS, 2050, South Africa.
E-mail address: seyiphilip@gmail.com (O.P. Oladijo).
0263-4368/$ see front matter 2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.ijrmhm.2014.01.009
Contents lists available at ScienceDirect
Int. Journal of Refractory Metals and Hard Materials
j our nal homepage: www. el sevi er . com/ l ocat e/ I JRMHM
17%Co coatings thermally sprayed onto metal substrates by means of
highenergy synchrotronX-rays. He foundthat the large strainmist be-
tween the coating and substrates emanated from the grit-blast surface
preparation of the surface, extending down to 0.5 mm, whereas the
residual strain over the gauge volumes employed were low, owing to
localised relaxation due to the microcracking.
X-ray diffraction was used to determine the residual stress in this
work. This technique was chosen because of the following reasons [13]:
It is non destructive, and the same specimens can be used for other
investigations, or repeated measurements can be performed.
It is phase distinctive, and capable of stress investigation in each
phase.
It has a moderate restriction on specimen dimensions and shape,
therefore, it can be used for measurements on different specimen
sizes and components.
The residual stress investigation of this research work had been re-
ported elsewhere [14], but inconclusive. This is due to the lattice spacing
{WC 202}, as well as singled spot considered for residual stresses deter-
mination on the large coated sample. Also, the correlation between the
residual stresses and abrasion resistance could not be nalised [16]
since the abrasive wheel irradiated area was larger than initial singled
spot. Thus, the objective of this work was to further explore the use of
X-ray diffraction to determine the residual stress associated with ther-
mally sprayed WC17Co coating on metal substrates. The work was
also done to investigate the inuence of residual strain on the abrasive
wear resistance, as well as coating properties (yield strength, surface
roughness).
2. Principles of X-rays diffraction
The principles governing the 2D detectors had been intensively
discussed elsewhere [14], whilst the principles concerning the sin
2

shall be discussed in details.


Residual stress magnitudes are determined through measurement
of changes in the material lattice spacing, d-spacing, due to the presence
of a stress. Based on the knowledge of the non-stressed lattice spacing,
any stresses present in the sample or material can be calculated using
the established sin
2
equation from Noyan [15]. In this method, a
colliminated X-ray beam of wavelength similarly to the interplanar
spacing is focussed onto a specimen and the number of X-rays diffracted
is counted as the angle between the X-ray detector and X-ray tube is
varied [16]. This allows a plot of diffracted intensity versus 2Theta to
be achieved. From these peaks, the lattice spacing, which varies from
stressed to non-stressed material can be determined using the Bragg
equation. Fig. 1 indicates the impinging and diffracted X-ray beam on
a magnied level. The angle is the angle between the surface normal
and the bisector of the incident and diffracted X-ray beam. It is also
the normal between the diffracted lattice planes and the sample's sur-
face. Fig. 1a) shows the sample orientated so that the diffracting lattice
planes are parallel to the surface. The compressive stress observed
would not affect this lattice spacing, as they are acting parallel to the
diffracting lattice planes. The sample has been rotated through a
known angle in Fig. 1b). The presence of compressive stress causes
the lattice spacing to be smaller than in the non-stressed state, which
can be measured through determination of the shift in the diffracted
intensity peaks. Once the shift is measured for at least two orientations
of y (the rotation of axis), then the lattice spacing, and hence residual
stresses, can be resolved.
Using the Bragg equation:
n 2dsin 1
where = wavelength (nm), n = constant, = Diffraction angle
(2 Theta in ).
The strain is obtained as the change in d-spacings between crystallo-
graphic planes and is given by:
d

d
o
do

1 v
E

sin
2

v
E

11

22
2
This predicts a linear variation of strain or interplanar spacing varia-
tion with sin
2
so that stress

can be obtained fromthe slope of a plot


of strain vs Sin
2
. The geometry for the biaxial stress is shown in Fig. 1.
where

is the stress component along the S

direction in the plane,


and is given by:


11
cos
2

12
sin
2

22
sin
2

spacing in the direction defined by and m ;


v Poisson ratio and E Youngs modulus Pa ;

Surface stress defined by the angle Pa


3
A graph of d

vs sin
2
is plotted with the stress value determined
from the slope [15]. This approach is called the sin
2
technique.
The slope m, is given by:
m 1 n =E

4
The parameters v/E and (1 + v)/E are generally known as S
1
and
1/2S
2
respectively, and are referred to as the X-ray elastic constants. A
linear behaviour of d

vs sin
2
is predicted by Eq. (2) [15].
3. Experimental procedure
A 2xxx Aluminium alloy, 304 L stainless steel, brass and super-invar
of 75 25 10 mm
3
in size were used as substrates. Wire EDM was
used to cut the larger coated samples into 30 25 10 mm
3
pieces
(i.e. small sample). The small samples were used for stress relaxation
investigations. The residual stress analysis was taken at three different
positions onthe surface of the coatedlarge samples (typical coatedsam-
ple used for abrasive wear tests discussed in previous publication [17]).
Fig. 1. Sample and laboratory coordinate systems [16]: a) = 0 and b) =i (sample is rotated through some known angle i), where D = X-ray detector, S = X-ray source, and N =
Surface normal.
69 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
The sample sizes and measuring positions are shown in Fig. 2. Micro-
structures investigation were done on 25 25 mm
2
sample cut from
the remaining procured substrate after the HVOF process, in order to
prevent stress induced in the material needed for residual stress
investigation.
Commercially available WC17 wt.%Co was used as a feedstock.
The powder was received with a nominal size distribution of (45 +
15 m) and was sprayed in this form. The surface of the substrate was
grit-blasted with alumna before deposition. Coatings of about 200 m
thickness were deposited by High Velocity Oxyl-fuel (HVOF) on 25
75 10 mm
3
top surfaces for four different substrates namely brass,
304 L stainless steel, super-invar and 2xxx aluminium alloy samples.
These substrates were selected because they have coefcients of ther-
mal expansion different from the coated material, and to each other to
assess the residual stress associated with the each coefcient of expan-
sion. Deposition parameters were the same on all coated samples, and
were as follows: 4 in. gun barrel; 380 mm spray distance; 0.0227 m
3
/h
fuel (kerosene) ow rate; and 56.6 m
3
/h oxygen ow rate, thus similar
to the parameters used elsewhere [14].
Residual stress measurements taken on coating surface were done on
an X-ray analyser with Co-K radiation, using the sin
2
method. The
anode settings were 40 kV and 40 mA. For the WC coating, the {112}
peak (reection at 123.62) that presented the highest diffracted intensi-
ties were used. All other peaks had signicant overlap with uorescence
peaks fromthe highenergy of the X-ray beam. The strainwere converted
to stress using bulk elastic constants S1 =3.247 10
7
MPa, 1/2S2 =
1.948 10
6
MPa and v = 0.20 [14]. Diffraction collection was done
using a two-dimensional High Star (Bruker AXS) detector. Data were
analysed using Leptos software version 6, as part of the Bruker AXS
suite of software for residual stress analysis. The full techniques
employed to determine the residual stress measurement is described
elsewhere [14]. Similarly, coating characterisation had been reported
elsewhere [6,17], which the results relevant to this research work listed
in Table 1.
The wear analysis was done using three abrasive wear tests on a
ASTM-G65 dry sand rubber wheel apparatus. An applied load of 25 N
and rotational wheel speed of 140 revs/min were employed on the
large coated samples. Each coated sample was abraded for 30 min. A
detailed description of the abrasive wear test and analysis is given else-
where [17] with the results listed in Table 2 for ease of reference.
The wear damage of the coating surface was further studied by
atomic force microscopy (AFM) using a Veeco Dimension 3100 and
XRD for qualitative analysis. While the state of abrasive silica sand
after post wear test was subsequently investigated using sieves analysis
method.
4. Results
The starting powder morphology and its cross-sectional part are
shown in Fig. 3(a and b). The typical SEM/EDX image of the thermal
spray WC17Co coatings showed that the light phase consisted of WC
grains and dark phase was the cobalt binder while the black portion is
the pore (Fig. 3(c)).
To understand the inuences of residual stress on the abrasive wear
resistance, the residual stress was determined at three different posi-
tions on large coated sample (typical sample used for wear test analysis
elsewhere [17]). These positions were chosen due to the abrasive wheel
irradiated area being larger than initial spot employed in the previous
published paper. The results are shown in Fig. 4, and the total average
values are given in Table 2. The residual stresses in WC phase over the
entire sample were compressive, but quite different in the different
positions, which implied that additional factors determined the stress
response on the large coated substrates.
Fig. 2. Sample geometry and measurement position for the strain analyses. Strain measurement was taken at positions 1, 2 and 5.
Table 1
Material characteristics [6,17].
Substrate CTE [10
6
/C] WC grain size (m) XRD results on coating Coating HV (GPa) Substrate HV5 (GPa)
Aluminium 2xxx series 23 0.150 0.01 WC, W
2
C, & Co 10.22 0.02 0.09 0.01
Brass 19 0.127 0.01 WC 10.04 0.01 1.40 0.01
304 L SS 17.3 0.185 0.02 WC 9.41 0.01 2.71 0.01
Super-invar 1.2 0.133 0.02 W
3
C, Co 7.91 0.01 1.53 0.01
CTE = Coefcient of thermal expansion.
70 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
Figs. 5 and 6 show the correlation that exists between wear loss,
yield strength and average residual stress determined on the large coat-
ed samples, which was remarkable.
The residual stress measured on the worn surfaces of the abrasive
coated samples reported elsewhere [17], and their results were com-
pared to the average residual stress on the large coated sample shown
in Fig. 7. The compressive residual stress values measured on the coated
sample increased after wear test.
XRD pattern of the worn coating surface are shown in Fig. 8. The
worn coating surfaces were largely dominated by WC peak as expected
with a minor W
3
O peak, except worn coated brass which had SiO
2
in
additions.
The AFMthree dimensional views of the worn surfaces of WC17Co
coatings were shown in Fig. 9. The images showed similar imperfection
that occurred on the worn surface after the abrasion. However, the wear
features were different among the substrates.
Fig. 10 shows the particle size distribution of the silica sand before
and after wear test. After post wear test, the sand showed slightly little
changes distribution.
In order to investigate the effect of stress relaxation in the coating.
The large coated sample cut by EDMwere re-examined by X-ray diffrac-
tion. The residual stress determined under the assumption of planar
stress conditions after cutting by wire EDMis showninFig. 11. After cut-
ting, the small coatedsamples size showeda general trend of decreasing
residual stress. The coated 2xxx aluminium alloy and super-invar sub-
strates consistently showed small changes. The coated 304 L stainless
steel changed from compressive stress to tensile stress, while the stress
of the coated brass sample was about a third of the large coated sample
before cutting (but still compressive).
5. Discussion
The morphology of the dried starting powder (Fig. 3a) and its cross
sectional images (Fig. 3b) was relatively simple. The powder showed a
dense structure with WC clusters cemented by cobalt, which were
both identied fromXRDand EDX. Microstructural analyses of the coat-
ing cross sections indicates that the typical thermal sprayed WCCo
coating, comprised equiaxed carbide particles distributed in the Co
matrix (Fig. 3c).
The experimental results of the residual stresses found on the large
as-sprayed coated samples (Fig. 4) were compressive stress but quite
different from each other. The compressive stress in coated 2xxx alu-
minium alloy and 304 L stainless steel increased, despite change the
sample measuring position. However, the compressive stress on the
surface of coated super-invar decreases with respect to the measuring
position. The coated brass showed a distinct difference between the re-
sults measured on its surface, which might have been due to the tex-
turing effect observed on its surface during measurement. Therefore,
the non-uniform measurements observed on the coatings with re-
spect to the measuring position might be due to the following
mechanism:
(1) porosity;
(2) irregular distribution of powder resulting from spray kinetic
[18];
(4) different grain size distribution;
(5) shape of the coated samples;
In addition, the rate of cooling can signicantly inuence the residu-
al stress or strain prole in the coating, which was not considered here,
because the cooling rates were assumed to be same.
The average residual stress (i.e. total) on the large coated substrate
samples are given in Table 2. The coated 2xxx aluminium sample had
the highest compressive stress, and the residual stresses of the coated
brass and 304 L stainless steel were in the middle, with the coated
super-invar having the lowest compressive stress value. The difference
in residual stress results measured on all the coated samples, despite
the same WC17Co powder being used as a feedstock was deduced to
be due to their differences in the coefcient of thermal expansion, inter-
action between coating and the substrate, as well as decarburisation
process, which had been discussed by the current author elsewhere
[6,14].
Many factors such as thermal history and deposition parameters
make it increasingly difcult to compare results of coating from other
sources [14]. Stokes and Looney [19] determined the residual stresses
in WCCo coatings by an analytical method similar to that of Clyne
et al. [20], and reported tensile stresses of 82 MPa and 15 MPa for coat-
ings of thicknesses of 200 and 600 m respectively. Ahmed et al. [10]
investigated the inuence of vacuum heat treatment on the residual
stress of thermal spray cermet coatings using neutron diffraction. They
Table 2
Yield strength, abrasive wear results and average residual stress (i.e. total) of the large coated samples.
Substrate Aluminium 2xxx series Brass 304 L stainless steel Super-invar
Coating yield strength (GPa) [17] 2.41 2.37 2.22 1.87
Wear mass loss (g) [17] 0.10 0.01 0.093 0.01 0.085 0.01 0.092 0.01
Average residual stress (MPa) 224.23 42.3 149.3 42.7 93.06 43.8 78.86 41.2
Surface roughness (nm) [6] 35 32 10 24
a b c
Fig. 3. SEM/BSE images showing (a & b) WCCo starting powder and its cross sectional view, showing WC clusters embedded in a cobalt matrix (c) coatings on the 304 L stainless steel
substrates, showing WC grains (light), cobalt binder (grey), and black pores.
71 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
observed the average values of the stress in the as-sprayed and heat-
treated coating layers to be 553 MPa and 492 MPa respectively.
The changes in the stress gradient after heat treatment, was related to
the compositional changes, caused by diffusion zones at the coating
substrate interface. The residual stresses of the samples investigated
were much lower than those of Ahmed et al. [10], which could be due
to the differences in feedstock powder, process, substrate sample and
surface roughness. The aws in the coating [21], localised microcracking
within the coatings, surface roughness, as well as surface stress relaxa-
tion might also have effects.
It is thought that differences in the employed techniques might con-
tribute to the lowvalues. Clyne et al. [20] highlightedsome difculties in
using X-ray measurement to determine the coating stress, including:
(1) Limitation of the range of , since large values would require the
incidence X-ray beam to penetrate appreciable thicknesses of
coating and/or substrate.
(2) The penetration depth of X-ray diffraction is proportional to the
surface roughness of as-sprayed coating only, as well as
(3) Error from variation in stress levels.
However, the greater penetration of neutrons offers several
advantages:
(1) Surface effects, such as deformation due to grinding or polishing,
or oxidation are avoided:
(2) Complete rotation of the sample is achieved, giving more thor-
ough sampling.
In general, the difference between the residual stress values ob-
served and those in the literature depend not only on the technique
and deposition parameters, but also on the method employed for
Fig. 4. Residual stresses taken at the three different positions on the large coated
substrates.
b
c
a
Fig. 5. Correlation between compressive the residual stress and (a) yield strength, (b) wear mass loss, (c) surface roughness of the WC17Co coatings on the different substrates.
72 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
deposition, which can have newphases, resulting in residual stresses as
the volume changes due to different coefcients of thermal expansion
[22].
There was a concern that the stresses were too small to have real
effect on the abrasion resistance and coating properties, although
Fig. 5 and Table 2 demonstrated strong correlations between yield
strength, residual stress, surface roughness and wear resistance of the
WC17Co coatings. The higher yield strength samples had higher resid-
ual stresses, showing that the residual stress was proportional to the
yield strength. Surface topography plays an important part in under-
standing the nature of coatings [23]. It is thought that the surfaces on
whichthermal spraying is done have to be chemically and mechanically
well-prepared for good adhesion. With an increased surface roughness,
the residual stresses increased as expected, thus the residual stress was
proportional to the surface roughness.
In order to investigate the inuence of residual stresses on abrasive
wear resistance of the WCCo coating, it is better to understand the
source of residual stresses in the abrasive wear measurement. It is
thought that the residual stresses in the wear process comes from two
main sources [24]: (1) residual stresses, resulting from friction and
wear process, that aid(tensile stresses) or retard (compressive stresses)
the wear process; (2) induced residual stresses of the required sign
(tensile or compressive) before the wear process or friction. Higher
wear resistance is accompanied by increased compressive residual
stresses, demonstrating that the abrasive wear resistance was propor-
tional to the compressive residual stress, similar to the work of Garba
[24]. However, the discrepancy of the as-coated super-invar was due
to partial stress relaxation, caused by the microcracking in the coating.
The wear resistance of all coating increased with an increasing yield
strength and surface roughness (Fig. 6).
In order to fully understand the stress state condition of the worn
coating surface after abrasive wear test, it is better to investigate the
state of residual stresses after the wear measurement (i.e. most abraded
part of the sample). Fig. 7 and Table 2 shows the residual stresses of as-
coated samples compared to after wear testing. The residual stresses
that were determined after wear test were compressive. It is thought
that the grinding of the coating by the abrasive sand under action of
force could have given rise to the plastic deformation and the introduc-
tion of compressive stresses. In addition, the compressive residual
stresses after the wear tests were higher than the as-coated samples
(i.e. before the wear test). This is expected as the surface of all the
worn samples had been altered by silica sand [17], as well as the inu-
ence of material shakedown. Stoica et al. [25] investigated the residual
stress and strain results of WC12Co coating in pre- and post tribologi-
cal test. He found that the residual stress before and after wear test were
compressive, with the residual stress value slightly higher at the end of
wear test due to inuence of material shakedown.
The XRD analyses of the worn coating surfaces (Fig. 8) showed new
phases as expected. All the worn coating surfaces comprises of WC peak
and W
3
O peak (minor), except worn coated brass which had SiO
2
in
additions. It is thought that W
3
O peaks could be formed as a result of
the interaction that existed between W (from the coating) and oxide
resulting from their wear and friction. While SiO
2
found on the worn
coated brass is in agreement with the SEM/EDX result. Comparison of
the XRD spectra of the as-sprayed coating [17] (with result listed in
Table 1) and the worn coating surface showed greater reduction in the
XRD intensity. This reduction might occur as a result of the residual
stresses generated from the interaction of abrasive wheel, silica sand
and the coating surface. Surprisingly, the initial secondary phases ob-
served on the as-sprayed coating (Table 1) could not be found on the
worn coating surface. It is thought that the secondary phases might
have been abraded or eroded during the interaction of the silica sand
on the coating. Although, the actual amount of the secondary phases
was not determined, but their peak were all minor which can be easily
get abraded by the sand particles.
Atomic force microscopy (AFM) was used to obtained more details
about the wear mechanisms, as its give qualitative information on the
binder and carbide elimination. Inspection of these images (see Fig. 9)
revealed that similar wear features were observed on the entire worn
coating surface. However, the degree of damage or evolution was
seemed different to one another. The worn coated 2xxx aluminium
sample suffered the greatest damage due to extensive matrix and
a b
Fig. 6. Correlation between wear mass loss and (a) coating yield strength, and (b) surface roughness of the WC17Co coatings on the different substrates.
Fig. 7. Residual stresses of the coated WC17Co samples, in the as-coated condition, and
after wear testing.
73 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
30 40 50 60 70 80 90 100 110 120 130 140
0
10000
20000
30000
40000
50000
60000
SiO
2
WC
WC
WC
WC
W
3
O
WC
WC
WC
WC
WC
WC WC
WC
WC
WC
304L SS
Super-Invar
Brass
2xxx Al
I
n
t
e
n
s
i
t
y
Position (2 Theta)
Fig. 8. XRD pattern of the worn coating surfaces.
a
b
c
d
Fig. 9. AFM Three dimensional view of the worn surfaces of WC17Co coatings on (a) 2xxx aluminium (b) brass (c) super-invar (d) 304 L stainless steel.
74 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
carbide removal. Fewer binder removal with strong standing carbide in
most area unaltered were observed on the worn coated brass and
super-invar, with extremely fewfound on the worn coated 304 L stain-
less steel. The worn coated 304 L stainless steel showed limited binder
removal with high carbide grains (upper area) unaltered, while 2xxx
aluminium reected the highest level of carbide grain and binder
removal. Thus in agreement with the SEM/EDX results discussed else-
where [17]. Guilemany et al. [26] reported that the excellent wetting
and adhesive characteristics of cobalt being binder make it difcult for
the carbide to be eliminated. Thus, a large quantity of binder material
elimination is necessary to begin removal of carbide. In addition, the dif-
ferences in the degree of damage onall four worn coating surfaces could
be due to their differences in substrate hardness, surface roughness, and
their physical properties that played a crucial role in cooling and solidi-
cation of the coating after HVOF process.
The particle size distribution of the abrasive silica sand used was
further studies by Mabotja [27]. This was done to verify its condition
after crushing the silica sand particle during the wear test. The results
(Fig. 10) showed no signicant changes of the particle sizes distribution
in comparison to before wear test. Thus, suggested that the silica sand
can still be used for more measurement, provided of the same proper-
ties to avoid any contamination.
Some understanding of the stress relaxation may be gained by
cutting the material into small sizes. After cutting the samples to pro-
mote stress relaxationby EDM(Fig. 11), there was little stress relaxation
in the coated 2xxx aluminium alloy, brass, and super-invar samples
which are still relevant during service (i.e. still benecial for an engi-
neering application). However, high stress relaxation was found in the
coated 304 L stainless steels, and it had changed to tensile residual
stress, which can leads to cracking and promote fatigue failure if the
magnitude exceeds the tensile strength of the coating [18]. A possible
explanation for this behaviour is that melting was observed to be the
primary material removal mechanism during EDM machining of the
sample. Thus, upon solidication, considerable thermal residual stresses
on the surface layer occurred which might inuence the performance of
the service.
It is thought that thermal treatment may affect the stress state of the
coating [12]. Thus a better knowledge of the material behaviour at high
temperature (i.e. heat treatment) is therefore needed in our future
study to improve the quality of the coating during service.
6. Conclusions
The following conclusions were drawn based from the determina-
tion of residual stress of WC7Co thermally sprayed onto different
metal substrates. The surface residual stresses determined by X-ray dif-
fraction are compressive and could be due to differences in coefcient of
thermal expansion and cooling rate. After cutting into small specimens
via wire EDM, the stress relaxation effect was high on coated brass, but
muchless on the other samples. There was correlation between residual
stresses, abrasive wear and coating properties. While AFM gives more
details information of the damage produced on worn coating surface.
Acknowledgements
The authors wishtoacknowledge the nancial andtechnical support
received from the Department of Science and Technology in South
Africa, National Research Foundation, the Nuclear Energy Corporation
of South Africa (NECSA), and the University of Witwatersrand. The
authors also wish to thank Tshepo Ntsoane and Ryno van der Merwe
from Necsa for assistance with XRD and SEM.
References
[1] Ahmed A. Rolling contact fatigue, analysis and prevention. ASM Handbook. ASM
International; 2003 94156 [Section 6E, 11].
[2] Osawa O, Itsukaichi T, Ahmed R. Inuence of substrate properties on the impact
resistance of WC cermet coatings. J Therm Spray Technol 2005;14(4):495501.
[3] Chuanxian C, Bingtain H, Huiling L. Plasma wear resistance of ceramic and cermet
coating materials. Thin Solid Films 1984;118:48593.
[4] de Villiers Lovelock HL. Powder/processing/structure relationships in WCCo
thermal spray coatings: a review of the published literature. J Therm Spray Technol
1998;7(3):35773.
[5] Sobolev VV, Guilemany JM, Nutting J. High velocity oxy-fuel spraying. UK: Maney
publishing; 2004.
[6] Oladijo OP, Venter AM, Cornish LA, Sacks N. PM2010 Powder Metallurgy World
Congress Proceedings, Florence, Italy; 2010. p. 18 [3(66)].
[7] Mantyla TA, Niemi KJ, Vuoristo P, Barbezat G, Nicola AR. Abrasion wear resistance of
tungsten carbide coatings prepared by various thermal spraying technique. In:
Eschnauer H, editor. Proceedings of 2nd Plasma-Technik Symposium, Lucerne,
Switzerland; 1991. p. 28797 [1, 1(1)].
[8] Liao H, Normand B, Coddet C. Inuence of coating microstructure on the abrasive
wear resistance of WC/Co cermet coatings. Surf Coat Technol 2000;124:23542.
[9] Verdon C, Karimi A, Martin J-K. Microstructural and analytical study of thermally
sprayed WCCo coatings in connection with their wear resistance. Mater Sci Eng A
1997;234236:7314.
[10] Ahmed R, Yu H, Edwards L, Santisteban JR. Inuence of vacuumheat treatment on the
residual stress of thermal spray cermet coating. Proceeding of the World Congress on
Engineering, vol. 11. 2007 [July 24, London, UK. ISBN 978-988-98671-2-6].
[11] Montay G, Cherouat A, Nussair A, Lu J. Residual stresses in coating technology.
J Mater Sci Technol 2004;20(1):814.
[12] Venter AM, Pirling T, Buslaps T, Oladijo OP, Steuwer A, Ntsoane TP, et al. Systematic
investigation of the residual strain associated with WCCo coatings thermal sprayed
onto metal substrates. Surf Coat Technol 2012;206(1920):401120.
[13] Matejicek J, Sampath S, Dubsky J. X-ray residual stress measurement in metallic and
ceramic plasma sprayed coatings. J Therm Spray Technol 1998;7(4):48996.
[14] Oladijo OP, Venter AM, Cornish LA, Sacks N. X-ray diffraction measurement of
residual stress in WCCo thermally sprayed onto metal substrates. Surf Coat Technol
2012;206:47259.
[15] Noyan IC, Cohen JB. Residual stress. New York: Springer Verlag; 1987.
[16] Stress-in-alloy forging. http://Sites.Google.com/site/Temfemguy/Research/
Residual.
[17] Oladijo OP, Sacks N, Cornish LA, Venter AM. Effect of substrate on the 3 body
abrasion wear of HVOF WC17 wt%Co coatings. Int J Refract Met Hard Mater
2012;35:28894.
[18] Kharlamov YA. Detonation spraying of protective coatings. Mater Sci Eng
1987;93:137.
Fig. 10. Particle size distribution of silica sand before and after wear tests.
Fig. 11. Residual stress comparison in large sample before and after cutting to medium
size.
75 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876
[19] Stoke J, Looney L. Residual stress in HVOF thermally sprayed thick deposits. Surf Coat
Technol 2004;18:1778.
[20] Clyne TW, Gill SC. Residual stress in thermally sprayed coatings and their effect on
interfacial adhesion. J Therm Spray Technol 1996;5(4):40118.
[21] Krawitz AD. The use of X-ray stress analysis for WC-base cermets. Mater Sci Eng
1985;75:2936.
[22] Pejryd L, Wigren J, Greving DJ, Shadley JR, Rybicki EF. Residual stresses as a factor in
the selection of tungsten carbide coating for a Jet engine application. J Therm Spray
Technol 1995;4(3):26874.
[23] Oladijo OP. Investigation of techniques for determining the residual stresses
in WC17Co thermal sprayed coatings and studying the effect of residual
stress on abrasion resistance. (Ph.D. Thesis) University of Witwatersrand;
2013.
[24] Garba II. Correlation between abrasive wear resistance and changes in structure and
residual stresses of steel. Tribol Lett 1998;5:2239.
[25] Stoica V, Ahmed R, Golshan M, Tobe S. Sliding wear evaluation of hot iso-
statically press thermal spray cermet coatings. J Therm Spray Technol 2004;
13(1):93107.
[26] Guilemany JM, Miguel JM, Vizcaino S, Climent F. Role of three-body abrasion wear in
the sliding wear behaviour of WCCo coatings obtained by thermal spraying. Surf
Coat Technol 2001;140:1416.
[27] Mabotja M. Research project report. University of the Witwatersrand; 2010.
76 O.P. Oladijo et al. / Int. Journal of Refractory Metals and Hard Materials 44 (2014) 6876

Você também pode gostar