You are on page 1of 11

Acta Materialia 51 (2003) 42674277

www.actamat-journals.com

Microhardness variation in relation to carbide development


in heat treated Cr3C2NiCr thermal spray coatings
S. Matthews, M. Hyland , B. James
The Department of Chemical and Materials Engineering, The University of Auckland, Private Bag 92019, Auckland, New Zealand
Received 22 January 2003; received in revised form 30 April 2003; accepted 5 May 2003

Abstract
Cr3C2NiCr thermal spray coatings have been extensively used to mitigate high temperature wear. During deposition
compositional degradation occurs through dissolution of the carbide phase into the matrix. High temperature exposure
leads to transformations in the microstructure, which influences the coating microhardness. While such developments
have been investigated in short-term trials, no systematic long-term investigations of the microhardness variation as a
function of microstructural development have been presented. In this work, high velocity sprayed Cr3C2NiCr coatings
were heat treated at 900 C for up to 60 days in air and argon. With treatment, matrix phase supersaturation was
reduced, while widespread carbide nucleation and growth generated an expansive carbide skeletal network. An initial
softening of the coatings occurred through matrix phase refinement, the subsequent hardness recovery was a function
of carbide development. Treatment in air generated further hardness increases as a result of internal oxidation.
2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
Keywords: Ageing; Isothermal heat treatment; Cermets; Hardness; Thermal spray coatings

1. Introduction
Thermally sprayed Cr3C2NiCr cermets have
been used extensively to mitigate wear and erosion
at high temperatures where WC based composites
are not suitable [1,2]. During thermal spraying,
substantial variations in the cermet composition
and microstructure are possible due to exposure to
the high temperature accelerating gas [37]. Dissolution of the carbide phase into the alloy matrix

Corresponding author. Tel.: +64-9-3737999; fax: +64-93737463.


E-mail address: m.hyland@auckland.ac.nz (M. Hyland).

occurs, reducing the carbide concentration and


generating marked variability in the matrix phase
composition. As a result carbon may be lost as CO
or CO2, altering the composition of the carbide
phase, while the matrix composition may vary
from regions dominated by Cr through to the Ni
rich starting composition [6,7]. Formation of
oxides can occur on the outer particle surface and
may potentially be distributed within the body of
the particle should widespread melting of the
matrix occur [8]. This widespread compositional
disorder is trapped within the metastable structure
as a result of rapid solidification upon deposition.
In addition, the microstructure of the supersaturated matrix phase varies with the spray conditions,

1359-6454/03/$30.00 2003 Acta Materialia Inc. Published by Elsevier Science Ltd. All rights reserved.
doi:10.1016/S1359-6454(03)00254-4

4268

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

potentially ranging from zones of amorphous


material [6,9,10] through to nano-crystaline [9,10]
and microcrystalline phases [7].
The complex metastable state of the as-sprayed
coating means that there is a large driving force for
microstructural and compositional transformations
when exposed to elevated temperatures. The most
notable change is the precipitation of fine carbides
[7,11,12] or possibly oxides [1,10] accompanied by
recrystalisation of the matrix. The mechanical
properties also change, in particular the microhardness which is commonly related to the performance
of carbide coatings under wear conditions [13].
There is no consistent trend with heat treatment;
reports of plasma coatings generally indicate an
increase in hardness, while HVOF coatings may
soften or harden depending on treatment temperature and atmosphere, or gun or powder type used
for deposition [35,9,10,14]. It is not surprising
then, that several different mechanisms have been
proposed correlating the variation in coating
microhardness with changes in microstructure.
Industrially, it is the long-term performance of
Cr3C2NiCr coatings at elevated temperature that
is significant. This has not been widely addressed
[1,2,15], particularly for high velocity thermal
spray coatings where no long-term (24 h) systematic studies of microstructural evolution at high
temperature have been presented. The purpose of
this study was to look at the long-term evolution
in microstructure and hardness and to better understand the relationship between the observed
changes in microstructure and hardness. The
response of high velocity air fuel (HVAF) and high
velocity oxygen fuel (HVOF) Cr3C2NiCr coatings
was assessed after exposure at 900 C for periods
of up to 60 days in air and argon. The variation
in microstructure was characterised and correlated
with the variation in microhardness, the response
varying in a manner unexpected from previously
presented work.

HVAF and GMA Microjet HVOF thermal spray


systems. Coating-only samples were prepared by
cutting away the substrate with a metallographic
cutting machine. The thin layer of remaining substrate material was removed from the samples
using a hot 10% nitric acid solution. Copper plates
in contact with the samples were used to accelerate
the steel dissolution. The samples were then
ground and polished. Analysis by X-ray diffraction
(XRD) and energy dispersive X-ray analysis (EDS)
indicated no detectable variation in composition as
a result of acid treatment. The sample dimensions
after polishing were 15 mm 7 mm, the thickness
of the HVAF samples ranging from 650850 m,
while those of the HVOF samples were 200300
m.
Heat treatment was conducted at 900 C in air
and under a positive pressure of argon. Samples
were removed after 2, 5, 10, 20, 30, 40 and 60
days. This temperature was used to accelerate the
microstructural development that occurs at lower
operating temperatures with long-term exposure,
without significantly altering the mechanism of
development. In all cases a tarnish layer had formed and was removed by grinding. This was to
expose the underlying heat treated material that
was not influenced by oxidation. Compositional
analysis was performed using XRD (Bruker D8
Advance, Cu source at 40 mA and 40 kV). Analysis on mounted and polished cross-sections was
carried out using back scattered electron (BSE)
scanning electron microscopy (SEM) (Philips
XL30S FEG). Vickers microhardness measurements (LECO M-400 Hardness tester) were conducted on the coating cross sections under a load
of 300 g for 5 s. An average hardness was calculated from 10 indents per specimen.

3. Results and discussion


3.1. Powder characterisation

2. Experimental procedure
A cermet powder of nominal composition
75%Cr3C225%NiCr (WOKA 2075-NiCr) was
sprayed on to mild steel substrates using Aerospray

The agglomerated and sintered WOKA 2075NiCr powder showed clear contrast between the
carbide and NiCr matrix phases (Fig. 1) indicative
of negligible carbide dissolution. This was
reflected in the XRD analysis (Fig. 2) where only

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

4269

Fig. 2. XRD spectra of the Cr3C2NiCr WOKA 2075-NiCr


powder and the as-sprayed HVOF and HVAF coatings.

Fig. 1. BSE cross sections of a WOKA 2075-NiCr powder


particle and the as-sprayed HVOF and HVAF coatings.

Cr3C2 and NiCr were identified. The carbide grains


were evenly distributed, forming interconnecting
networks within each powder particle.
3.2. As-sprayed coatings
HVOF sprayed Cr3C2NiCr coatings have been
extensively investigated, the properties and appearance of the coatings in this work within the ranges
reported by others. BSE micrographs showed that
extensive carbide dissolution in the matrix

occurred, reducing the concentration of carbides


and generating widespread greyscale variation
indicative of marked variation in the matrix composition (Fig. 1). Retained carbide grains were
found predominantly in the centre of the larger
splats. Dissolution of carbide into the matrix and
the consequent super saturation of the matrix with
Cr and C gave rise to a large amorphous background on the low angle side of the NiCr peak in
the XRD scan (Fig. 2). The only detectable crystalline phases were Cr3C2 and NiCr: while small
amounts of Cr7C3 may have formed, its concentration was to low be distinguished from other
overlapping peaks. The splats were long and thin,
and often bounded by oxide stringers, indicating
that the majority of the powder particles were
molten at impact.
In marked contrast to the HVOF coating, a high

4270

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

carbide content was retained in the HVAF coatings, the isolated carbide grains being homogeneously distributed. The alloy phase showed significantly lower degrees of greyscale variation than
in the HVOF coating (Fig. 1). The lower degree
of carbide dissolution is also apparent in the XRD
spectra (Fig. 2). Only Cr3C2 and NiCr were
detected, the spectra of the latter dominated by narrow crystalline peaks with smaller amorphous
background. A striking feature of this coating was
the lack of visible splat structure. No inter-splat
oxides were evident and the lack of wide spread
carbide dissolution eliminated the matrix phase
contrast of the previous coating. There are few
published studies of HVAF Cr3C2NiCr coatings
for comparison, but the general features described
herelow degree of carbide degradation and
homogeneous nature of the coating, mirror features
seen in WCCo and WCCoCr coatings sprayed
by HVAF [1618].
3.3. Heat treatment of the coating samples
It is evident from the XRD spectra of the heat
treated HVAF and HVOF coatings (Fig. 3) that
microstructural
transformations
were
well
advanced after 2 days of exposure. The amorphous
regions of the as-sprayed coating spectra had disappeared, with narrow, intense NiCr peaks dominating the matrix phase spectra. The carbide peaks
were also more defined, and BSE images showed
that fine precipitates had formed. Similar changes
have been documented by others [35,7] and can
take place within 1 h at 800 C [3]. Image analysis
of the BSE images indicated that the carbide volume fraction rapidly increased from approximately
36% and 67% in the HVOF and HVAF as-sprayed
coatings, respectively, to a value of approximately
75 vol.% (Fig. 4). This remained stable over the 60
days of treatment for both the HVAF and HVOF
coatings under air and argon. The carbide itself
was not significantly oxidised. Quantitative analysis of the carbide XRD spectra in terms of the peak
area, FWHM, and centre of gravity of the primary
peaks indicated that Cr3C2 remained the dominant
carbide phase and was not oxidised to Cr7C3 in the
HVAF coatings, although a small amount may
have formed in the HVOF coatings. Cr2O3 was

Fig. 3. XRD spectra of the HVOF and HVAF coatings heat


treated for 2 days at 900 C.

Fig. 4. Volume percent carbide in the HVAF and HVOF coatings as a function of heat treatment under argon.

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

4271

detected in some of the XRD spectra, the low


intensity, noisy peaks indicating that this phase
was present in very low concentrations. This evidence indicates that the fine precipitates which formed during heat treatment were most likely Cr3C2
and not Cr2O3 as suggested by others [1,2,9,10].
3.4. Carbide precipitation and growth during
heat treatment
For the HVAF coating, precipitation of fine carbides occurred throughout the coating during the
initial 2 days of exposure, primarily in regions of
high carbide density (Fig. 5). This reflected the low
degree of particle heating in-flight as the dissolved
Cr and C only diffused short distances from the
parent carbides. Preferential nucleation occurred
on those carbides retained during spraying. Where
isolated precipitates occurred they tended to
coalesce and form larger carbides of complex morphologies ranging from blocky agglomerates
through to thin spidery networks.
Continued carbide growth occurred out to 5 days
with the average nucleated carbide size substantially larger than at the shorter exposure time. The
reduction in the number of such grains, and their
more rounded appearance suggested growth via
Ostwald ripening [19]. Widespread coalescence
and bridging occurred with an evident macroscopic
carbide skeletal network beginning to form.
Beyond 5 days development of the carbide structure began to slow. Growth occurred primarily
through strengthening of the inter-carbide bridging
networks and continued carbide coalescence, generating large single phase regions of complex morphology. Within this network the grains themselves
appeared more spherical. Beyond 30 days no significant changes were evident. The development of
the carbide skeletal network had occurred to the
extent where the carbide phase was largely continuous, surrounding pockets of matrix phase.
Treatment in air led to the same carbide phase
transformation with time as were noted in argon.
Oxides phases were evident in higher concentrations. Somewhat unexpectedly they did not form
inter-splat oxide stringers but were noted to occur
as small isolated pockets.
In the HVOF coatings, after 2 days treatment

Fig. 5. BSE images of the HVAF microstructure in the assprayed condition and after 2, 20 and 60 days of treatment under
argon at 900 C.

4272

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

under argon the extensive variation in the assprayed matrix phase greyscale contrast was significantly reduced (Fig. 6). Widespread carbide
precipitation occurred, the average size smaller
than in the HVAF coatings. Unlike the HVAF
coatings, carbide nucleation occurred primarily in
the carbide free zones where the greatest degree
of dissolution had occurred during spraying. The
variation in the magnitude of dissolution generated
a variety of carbide distributions and morphologies
with heat treatment. Three generalised regions
were evident (Fig. 7). Region 1 occurred in the
centre of the largest splats in pools of relatively
unaffected material. Carbide dissolution was minimal and the subsequent carbide nucleation mirrored that of the HVAF coating. Region 2 occurred
midway between the core and periphery of the
largest splats, or within those splats where widespread carbide dissolution had occurred. Spongelike agglomerates formed via coalescence of large
numbers of nucleated carbides. These were
expansive in nature, occupying significant regions
of the splat, but at the edges of the splat followed
the splat contour. At the splat boundary, Region 3,
the carbides formed long thin stringers that in three
dimensions may have formed plate-like morphologies.
Growth of the individual carbide grains led to
coarsening of the sponge-like agglomerates, generating a more porous appearance to these features.
This occurred most rapidly away from the splat
peripheries leading to widespread coalescence and
bridging. While the peripheral plate-like carbides
were still evident after 5 days they were beginning
to break up into more isolated grains. After 1020
days the structure developed to a similar state as
the HVAF coating. Growth and spheriodisation of
the individual grains occurred, while coalescence
and bridging between the grains led to the development of a widespread skeletal network. Such development of the carbide phase began to overshadow
the preferential orientation of the carbides in
relation to the splat shape. Beyond 30 days, growth
slowed in a similar manner to the HVAF coating.
After 60 days both the HVOF and HVAF coatings
were comparable, the HVOF carbides having a
slightly wider size distribution and more distorted
morphologies. It was evident, however, that despite

Fig. 6. BSE images of the HVOF microstructure in the assprayed condition and after 2, 20 and 60 days of treatment in
air at 900 C.

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

Fig. 7. Schematic illustration of the three generalised regions


of carbide development within a splat during heat treatment of
the HVOF coating.

the marked variation in the as-sprayed condition,


both coating systems tended towards the same
microstructure with heat treatment.
Treatment in air had no obvious influence on the
carbide nucleation or growth processes. The presence of oxide in the as-sprayed HVOF coatings
complicated the analysis of oxide ingress, however, the pockets of oxide and thickening of the
oxide stringers appeared more frequent in the air
treated samples.
3.5. Microhardness of Cr3C2NiCr coatings as a
function of heat treatment
The same general trend in microhardness
occurred for both the HVAF and HVOF coatings
in air and argon (Fig. 8). A rapid reduction in hardness occurred over the initial period of heat treatment. With continued exposure the hardness
quickly increased, reaching stable values after 20
30 days of treatment. In all instances the HVOF

Fig. 8. Vickers microhardness values of the HVAF and HVOF


coatings treated in air and argon at 900 C.

4273

coatings had higher hardness values, although the


initial period of softening was more dramatic and
took longer to recover in these coatings. Treatment
in air generated consistently harder coatings than
those treated in argon for both coating systems.
Because of the microstructural complexity of
thermal spray coatings, a number of hardness
mechanisms are likely to be responsible for the
changes in hardness with heat treatment and the
hardness differences between coatings sprayed by
HVAF and HVOF as shown in Fig. 8. For carbide
cermets, WC in particular, the model of Lee and
Gurland [20,21] has commonly been applied, Eq.
(1),
HC HWCVWCC HM(1VWCC)

(1)

In this model the cermet hardness (HC) is based on


the hardness and volume fraction of the hard phase
(HWC and VWC) and the hardness of the matrix
phase (HM). The contiguity (C), defined by Gurland [20,21], describes the percentage surface area
of a carbide grain in contact with other carbide
grains and accounts for the microstructural influence of the carbide morphology. The hardness of
the two phases is microstructure dependent. Carbide hardness is a function of grain size, smaller
grains giving higher hardness values [21,22]. Constraint of the matrix between the carbides elevates
the effective hardness value of this phase, higher
hardness values achieved with smaller inter-carbide spacing [21,22]. With regard to the overall
cermet hardness, higher degrees of contiguity generates harder material [20]. In addition, for thermally sprayed cermet coatings, splat-splat bonding,
porosity and oxide formation will also affect the
overall hardness.
The as-sprayed HVAF and HVOF coatings demonstrate this complexity. Despite the significantly
higher carbide content of the HVAF as-sprayed
coating, its hardness is lower than the HVOF assprayed coating. This is in contrast to the results
of Zimmermann and Kreye [3] who showed a correlation between higher retained carbide content
(as measured by C content of the coating) and
higher hardness in Cr3C2NiCr coatings sprayed by
three HVOF techniques. However, in their work
other influencing factors such as porosity were not
taken into account. In the present study, the coat-

4274

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

ings have similar levels of porosity (3%), the difference in hardness in the as-sprayed coatings is
likely to be due to differences in matrix hardness
the matrix of the HVOF coating will be significantly harder due to solid solution strengthening
generated by more extensive carbide dissolution
during spraying. When Cr3C2 dissolves into the
matrix the dissolved Cr and C distort the Ni lattice,
inhibiting dislocations motion [23]. The structural
disorder caused by greater amounts of dissolved Cr
and C is shown in the larger XRD peak widths for
the matrix phase in the HVOF compared to the
HVAF (Fig. 2). The extent of binder phase
strengthening is dependent on the deposition parameters, being more influential in the higher temperature techniques where extensive carbide dissolution means that the coatings act more as
dispersion strengthened alloys than carbide based
composites. The higher oxide content of the HVOF
coating could also affect its hardness, however, the
role of oxides is ambiguous. On the one hand, their
higher hardness will increase the overall hardness,
on the other, their presence as stringers may reduce
splat/splat bonding, lowering the measured hardness [4].
Following heat treatment in argon, both the
HVAF and HVOF coatings experience a decrease
in hardness with gradual recovery and strengthening after 2 or 5 days, respectively. Other
researchers have also recorded decreases in hardness for HVOF coatings using similar powder
types and similar treatment regimesin vacuum at
temperatures between 800 and 900 C, 2 days
[35]. In these studies the treatments were shortterm, so the recovery in hardness was not documented.
The changes in hardness with heat treatment
were accompanied by concurrent changes in
microstructurenamely the precipitation of fine
carbides and their subsequent growth; and the
recrystalisation of the matrix. Recrystalisation of
the matrix was evident in the reduced matrix contrast in the BSE images and in the decrease in dspacing of the Ni peak due to a reduction in the
amount of dissolved Cr and C. As indicated by the
narrowing of the matrix phase XRD peaks, recovery, recrystalisation and grain growth occur rapidly, reducing the dislocation density, increasing

the grain size and relieving any internal stresses


within the splats. These transformations lead to
matrix phase softening and are thought to account
for the initial reduction in coating hardness over
the first 25 days of exposure.
At the same time, nucleation of the carbide
phase takes place, which was shown to recover to
a stable concentration within the first 2 days of
treatment. The increase in carbide content is
thought to contribute to the recovery in hardness
with extended exposure, a mechanism favoured by
several authors [3,5]. Zimmermann and Kreye [3]
heat treated JP5000, Top Gun and Jet Kote HVOF
Cr3C2NiCr coatings at temperatures up to 800 C
for 1 h. The JP5000 coating showed the highest
hardness which slowly decreased at temperatures
approaching 800 C. The other coatings exhibited
peaks in hardness at 500 and 600 C, respectively,
also before softening. A similar treatment by
Otsubo et al. [5] observed stable hardness values
for JP5000 Cr3C2NiCr coatings but increasing
hardness values for a PlazJet III250 plasma coating. Both research groups attributed the increase in
hardness to carbide precipitation. Peaks in hardness
were proposed to result from an optimum carbide
size being achieved, while over-aging of the precipitates accounted for the softening in the higher
temperature trials [5]. The stable response of the
JP5000 coating was attributed to the minimal
degree of carbide dissolution in-flight and hence
the limited degree of carbide precipitation.
Carbide nucleation is also thought to increase
the coating hardness in the current work, but the
mechanism is different. Initially, carbide precipitation results in isolated or agglomerated groups of
grains. While these are small they effectively act
as dispersion strengthening particulates to harden
the matrix phase. Significant hardening of the overall cermet, however, only occurs with further
development of the carbide phase. The volume
fraction of the matrix and carbide phases were
shown to be constant after 2 days and hence,
according to Eq. (1), it is the variation in microstructure and contiguity that control the hardness
with extended exposure. In the HVAF coating this
occurs through bridging and coalescence of the
carbide grains resulting from localised nucleation
of carbides within the regions of high retained car-

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

bide density. Even where carbide interaction was


yet to develop significantly, the close proximity of
the hard particles generated high localised matrix
hardness. The homogeneous carbide distribution
ensured that such development occurred evenly
throughout the coating. In contrast carbide
nucleation in the HVOF coating occurred in carbide free zones, generating expansive sponge-like
agglomerates. Within such features the small carbide size and small matrix mean free path between
the carbides suggests that, as an entity, such clusters act effectively as larger reinforcing particles
with a hardness tending towards that of the larger
carbide grains. After 1020 days exposure, there
was a transition in both coatings from the growth
of individual grains to the development of a widespread carbide network. The larger carbides formed through agglomeration and coalescence
increase the volume of material over which the
load is applied, thereby increasing the resistance to
penetration. The earlier time at which widespread
carbide interaction occurs is thought to account in
part for the more rapid hardening response of the
HVAF coatings during the early stages of
exposure. With long-term exposure the rate of carbide development decreased, a trend reflected in
the stable hardness achieved after 30 days.
In addition to the microstructural changes the
physical splat structure of the coating also changes
with heat treatment. Diffusion results in solid state
sintering and the formation of physical bonds
between the mechanically interlocked splats. Such
growth was believed to account for the reduction
in porosity and splat boundaries of thermal spray
coatings with extended heat treatment at 800 and
900 C by Tobe et al. [15]. The resulting increase
in homogeneity of the coating structure also contributes to higher hardness values with long-term
exposure.
3.6. Influence of the treatment environment
Heat treatment of the coatings in air generated
higher average hardness values in both the HVAF
and HVOF coatings, but otherwise mirrored the
argon heat treatments. This differs from previous
studies, which showed an increase in hardness at
treatment times where the present coatings are still

4275

softening. It is likely that the overall increase in


hardness compared to argon treated samples is due
to oxides in the matrix. Oxidative hardening was
also suggested by Sahoo and Raghuraman [1,2],
He et al. [9,10] and Fagoaga et al. [14] in heat
treatment of Cr3C2 based coatings from blended
powders. The rise in hardness within the first 24 h
of exposure was attributed to oxidation of the carbides Cr3C2 and Cr7C3 directly to Cr2O3 without
any decarburisation steps. While in the work of
Sahoo and Raghuraman [1,2] this applied to large
carbide splats, a different mechanism was highlighted by He et al. [9,10] for nano-structured coatings. To some degree these responses may be
dependent upon the use of blended powders, particularly in regard to the coating structure.
In the present study using sintered and agglomerated powders, it is more likely that the matrix,
rather than the carbide is preferentially oxidised.
The higher degree of carbide dissolution generated
with the use of agglomerated and sintered powder
and the finer carbide size means that the binder is
more susceptible to oxidation. Hardening by this
mechanism would be more influential in the HVOF
coatings of this work due to the thinner sample
thickness and the greater splat surface area per
depth of coating resulting from the thinner nature
of the splats. As highlighted in the results of Sahoo
and Raghuraman [2] the influence of oxide formation occurs within the first 24 h as a protective
oxide layer forms with longer periods of exposure
[24]. As such this suggests that the influence of
internal oxidation occurred prior to the 2 day data
values, its effect remaining as a fixed increment of
higher hardness over the remaining trials.
4. Conclusions
In this work, the microhardness response of
HVAF and HVOF Cr3C2NiCr coatings was
assessed following treatment at 900 C for periods
of up to 60 days in air and argon. Initially, all
samples showed a drop in hardness. This was attributed to the reduction in strengthening mechanisms of the matrix phase with heat treatment. Of
those considered solid solution strengthening generated by carbide dissolution was thought to be the
most dominant.

4276

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

Hardness recovery and achievement of stable


hardness values after 30 days resulted from the precipitation and development of the carbide phase.
The air treated samples of both the HVAF and
HVOF coatings were consistently harder than
those treated in argon as a result of internal oxidation.
The HVOF coatings were harder than the HVAF
coatings, possibly as a result of the matrix strengthening through carbide dissolution or higher assprayed oxide content.

[7]

[8]

[9]
[10]

Acknowledgements
[11]

The authors gratefully acknowledge the assistance of WOKA, Metal Spray Suppliers (NZ) Ltd,
and Holster Engineering (NZ) Ltd for supplying
the powder and coatings for this work. The financial assistance provided by Material Performance
Technologies (NZ) and The University of Auckland is greatly appreciated.

References
[1] Sahoo P, Raghuraman R. Chromium carbide reinforced
composite coatings for high temperature hard-coat applications. In: Berndt CC, Bernecki TF, editors. Thermal
spray: research, design and applications. Materials Park,
OH, USA: ASM International; 1993. p. 40510.
[2] Sahoo P, Raghuraman R. High temperature chromium carbide reinforced metal matrix composite coatings for turbomachinery applications. In: TS93: Thermal Spraying
Conference (Thermische Spritzkonferenz). Deutscher Verlag fur Schweisstechnik DVS-Verlag GmbH; 1993. p.
296300.
[3] Zimmermann S, Kreye H. Chromium carbide coatings
produced with various HVOF spray systems. In: Berndt
CC, editor. Thermal spray: practical solutions for engineering problems. Materials Park, OH, USA: ASM International; 1996. p. 14752.
[4] Tomita T et al. Mechanisms of high hardness in Cr3C2
NiCr cermet coatings formed by vacuum plasma spraying.
In: Berndt CC, Khor KA, Lugscheider EF, editors. Thermal spray 2001: new surfaces for a new millennium.
Materials Park, OH, USA: ASM International; 2001. p.
699704.
[5] Otsubo F et al. Properties of Cr3C2NiCr cermet coating
sprayed by high power plasma and high velocity oxy-fuel
processes. J Therm Spray Technol 2000;9(4):499504.
[6] Guilemany JM, Calero JA. Structural characterisation of

[12]

[13]

[14]

[15]

[16]

[17]

[18]

[19]

[20]

[21]

chromium carbidenickel chromium coatings obtained by


HVOF-spraying. In: Berndt CC, editor. Thermal spray: a
united forum for scientific and technological advances.
Materials Park, OH, USA: ASM International; 1997. p.
71721.
Guilemany JM, Nutting J, Llorca-Isern N. Microstructural
examination of HVOF chromium carbide coatings for high
temperature applications. J Therm Spray Technol
1996;5(4):4839.
Sobolev VV, Guilemany JM. Effect of oxidation on droplet flattening and splatsubstrate interaction in thermal
spraying. J Therm Spray Technol 1999;8(4):52330.
He J et al. Thermal stability of nanostructured Cr3C2NiCr
coatings. J Therm Spray Technol 2001;10(2):293300.
He J, Lavernia EJ. Precipitation phenomenon in nanostructured Cr3C2NiCr coatings. Mater Sci Eng A
2001;301:6979.
Hidalgo V, Varela F, Martinez S. Characterisation and
high temperature behaviour of Cr3C2NiCr plasma
sprayed coatings. In: Lugscheider E, Kammer RA, editors.
Tagungsband Conference Proceedings. Germany: DVS
Verlag GmbH; 1999. p. 6836.
Taylor TA. Phase stability of chromecarbide NiCr coatings in low oxygen environments. J Vac Sci Technol
1975;12(4):7904.
Wayne SF, Sampath S. Structure/property relationships in
sintered and thermally sprayed WCCo. J Therm Spray
Technol 1992;1(4):30716.
Fagoaga I et al. Multilayer coatings by continuous detonation system spray technique. Thin Solid Films
1998;317:25965.
Tobe S et al. High temperature corrosion resistance of
newly developed Cr-based alloy coatings. In: Lugscheider
E, Kammer RA, editors. Tagungsband Conference Proceedings. Germany: DVS Verlag GmbH; 1999. p. 296
300.
Akimoto K, Horie Y. Study of HVAF WC-cermet coatings. In: Ohmori A, editor. Thermal spraying: current
status and future trends. Osaka, Japan: High Temperature
Society of Japan; 1995. p. 3136.
Jacobs L, Hyland MM, Bonte MD. Comparative study of
WC-cermet coatings sprayed via the HVOF and HVAF
processes. J Therm Spray Technol 1998;7(2):2138.
Jacobs L, Hyland MM, Bonte MD. Study of the influence
of microstructural properties on the sliding-wear behaviour of HVOF and HVAF sprayed WC-cermet coatings. J
Therm Spray Technol 1999;8(1):12532.
Martin JW. Micromechanisms in particle-hardened alloys.
In: Cahn RW, Thompson MW, Ward IM, editors. Cambridge solid state series. Cambridge: Cambridge University Press; 1980.
Lee HC, Gurland J. Hardness and deformation of
cemented tungsten carbide. Mater Sci Eng 1978;33:125
33.
Gurland J. Application of quantitative microscopy to
cemented carbides. In: McCall JL, editor. Practical appli-

S. Matthews et al. / Acta Materialia 51 (2003) 42674277

cations of quantitative metallography. Philadelphia: American Society for Testing and Materials; 1984. p. 6584.
[22] Engqvist H, Jacobson S, Axen N. A model for the hardness of cemented carbides. Wear 2002;252:38493.
[23] Felbeck DK. Introduction to strengthening mechanisms.
In: Dorn J, editor. Prentice-Hall series in materials science.
New Jersey: Prentice Hall; 1968.

4277

[24] Matthews S et al. Isothermal oxidation of Cr3C2NiCr


coatings sprayed by high velocity techniques. In: Lugscheider E, Berndt CC, editors. International Thermal Spray
Conference. Germany: DVS Verlag GmbH; 2002. p.
698704.

You might also like