Surface & Coatings Technology: Sciencedirect

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Surface & Coatings Technology 371 (2019) 322–332

Contents lists available at ScienceDirect

Surface & Coatings Technology


journal homepage: www.elsevier.com/locate/surfcoat

Tribological and high-temperature mechanical characterization of cold T


sprayed and PTA-deposited Stellite coatings
L. Baiamontea,b, , M. Tuluic, C. Bartulia,b, D. Marinia,b, A. Marinoc, F. Menchettic, R. Pileggic,

G. Pulcia,b, F. Marraa,b
a
Dept. Of Chemical Engineering, Materials and Environment, Sapienza University of Rome, Italy
b
INSTM Reference Laboratory for Engineering of Surface Treatments, Rome, Italy
c
Centro Sviluppo Materiali S.p.A., Lamezia Terme (CZ) Branch Office, Italy

ARTICLE INFO ABSTRACT

Keywords: Stellite 6 is a cobalt-base alloy commonly used for coatings useful in wear and corrosion protection of several
Cold gas spray components over a wide range of temperatures.
PTA Stellite coatings can be obtained by many deposition techniques: plasma transferred arc (PTA) is certainly an
Stellite 6 optimal choice that gives thick, dense coatings, but thermal spray techniques like high velocity oxy-fuel (HVOF)
Cobalt-base alloy
are also widely used to deposit such materials.
High temperature mechanical properties
The present study investigates wear and mechanical properties of Stellite 6 deposited by cold spray in
Tribology
comparison to PTA-deposited Stellite 6, aiming to obtain a very dense coating without the dilution zone typically
resulting from PTA weld surfaces.
Tribological behavior of cold sprayed Stellite 6 deposited on AISI 304 stainless steel was evaluated by block-
on-ring wear test; friction coefficients and mass and volume variations were analyzed, followed by SEM-EDS on
wear tracks.
Mechanical properties were evaluated by four-point bending test from room temperature up to 750 °C, and by
dynamic indentation. Vickers microhardness and porosity were also evaluated.

1. Introduction the microstructure of the coating; moreover, the equipment needed –


and thus the associated costs - differs from system to system. LPPS, for
Stellite is a family of cobalt-based alloys known for their out- instance, requires expensive vacuum plants, while other processes give
standing anti-wear and anti-corrosion applications over a wide range of oxidized or diluted coatings. Specifically, PTA-deposited Stellite, while
temperatures. [1,2] Chemical composition dramatically affects both the providing an adherent and coherent coating, suffers from the presence
properties and the microstructure of the Stellite alloys [3]; specifically, of a dilution zone that dramatically affects the microstructure and,
the carbon content has a crucial role because the carbide volume consequently, microhardness and in turn the tribological performance
fraction influences many key features (e.g., microhardness, ductility, of the coating [7].
wear resistance) of the alloy [4]. While HVOF is a suitable alternative to deposit the cobalt-based
Stellite 6 is one of the most common alloy of the Stellite family; its alloy, the need for a more cost-effective technology and an oxide-free
nominal composition is 12.6 wt% (Cr0.80Co0.15W0.05)7C3 carbides in a coating led to an increasing interest towards cold gas spraying (CGS)
Co-23Cr-4.5 W solid solution alloy [5,6]. This coating can be obtained deposition system [8] to obtain thick, dense, non-oxidized and non-
through various deposition techniques: plasma transferred arc (PTA); diluted coatings. CGS is based on plastic deformation of the powder
thermal spray methods such as low-pressure plasma spray (LPPS), high particles during impact on the substrate, without reaching neither total
velocity oxy-fuel (HVOF), cold spray; laser cladding, are all suitable nor partial melting of the feedstock materials [9].
processes for depositing cobalt-based alloys. The present paper aims to provide a preliminary comparison be-
Each deposition technique obviously affects the final properties and tween cold sprayed and PTA-deposited Stellite 6 alloy in terms of wear

Corresponding author at: Dept. Of Chemical Engineering, Materials and Environment, Sapienza University of Rome, Via Eudossiana 18, 00184 Rome, Italy.

E-mail addresses: lidia.baiamonte@uniroma1.it (L. Baiamonte), mario.tului@rina.org (M. Tului), cecilia.bartuli@uniroma1.it (C. Bartuli),
danilo.marini@uniroma1.it (D. Marini), angelo.marino@rina.org (A. Marino), fernando.menchetti@rina.org (F. Menchetti), rosanna.pileggi@rina.org (R. Pileggi),
giovanni.pulci@uniroma1.it (G. Pulci), francesco.marra@uniroma1.it (F. Marra).

https://doi.org/10.1016/j.surfcoat.2019.04.032
Received 31 August 2018; Received in revised form 13 March 2019; Accepted 8 April 2019
Available online 09 April 2019
0257-8972/ © 2019 Elsevier B.V. All rights reserved.
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Table 1 evaluated as the slope of the unloading curve at the maximum load
Cold spray deposition parameters. applied. Once obtained the reduced elastic modulus, the Young's
Gas temperature [°C] 970 modulus of the tested sample can be extracted from Eq. (2):
1
Pressure [bar] 40 1 1 2
s 1 i
2
Stand-off [mm] 20 = + ,
Scanning velocity [mm/s] 300
Er Es Ei (2)

with Es, Ei, νs and νi being the Young's moduli and the Poisson's
Table 2 ratios of specimen and indenter, respectively.
PTA deposition parameters. A Poisson ratio of 0.28 was selected.
Vickers microhardness measurements (ASTM E-384-89) [14] were
Current [A] 130
carried out on polished crossed section of the heat-treated samples,
Voltage [V] 30 with a Leica VMHT Microhardness Tester (Leica Microsystems GmbH,
Feed rate [rpm] 12 Germany); a 0.49 N (50 gf) load was reached in 10 s and then main-
Stand-off distance [mm] 8 tained for 15 s.
Scanning velocity [mm/s] 10
Four-point bending tests in air (ASTM C1161-18 for room tem-
perature and C1211-18 for high temperatures) [15,16] were performed
behavior and mechanical properties, in order to understand whether on freestanding samples upon removing the steel substrates by a cutting
and to what extent the presence or absence of a dilution zone and of and grinding procedure performed after the depositions. The selected
oxidized particles actually affect the cobalt-base alloy. test temperatures were: room temperature, 300 °C, 500 °C, 650 °C and
750 °C.
As prescribed by the related standards, a 3 N pre-load was applied
2. Materials and methods before the test. Tests were carried out using a Zwick-Roell Z2.5 testing
machine (Zwick GmbH, Ulm, Germany, Fig. 1a) equipped with a fur-
2.1. Materials nace (Maytec GmbH, Germany), a silicon carbide fully-articulated
flexure device (Fig. 1b), and a 3-point-contact extensometer specially
Cold sprayed samples were obtained from a Stellite atomized designed to obtain the curvature of the sample and so its deformation.
powder (Diamalloy 4060 NS, Oerlikon-Metco, Switzerland) was de- Stress-strain curves were elaborated by Zwick-Roell TestXpert II soft-
posited onto AISI 304 stainless steel substrates with an Impact 5/11 ware: the elastic modulus was calculated in the elastic interval between
Cold Gas Spray (Impact Innovations GmbH, Haun, Germany) facility 20 and 40 MPa [17,18].
installed at Centro Sviluppo Materiali (Rina Consulting – C.S.M. SpA, Microstructural analysis was carried on chemically etched samples
Lamezia Terme unit, Italy). after four-point bending tests, to investigate possible variations in mi-
Table 1 shows the deposition parameters. crostructure of both cold-sprayed and PTA-deposited coatings after high
PTA coatings were deposited using a Stellite 6 alloy (Kennametal, temperature tests. The samples were etched with aqua regia (30 ml HCl
Latrobe, PA, USA) on AISI 304 stainless steel substrates, with argon as 37% + 10 ml HNO3 65%). [19] Energy Dispersive X-ray Spectrometry
shielding gas. The deposition parameters are reported in Table 2. (EDS) analysis were performed on etched samples from both type of
deposited Stellite 6, using a scanning electron microscopy (Philips SEM
3. Methods XL40) device assisted by EDAX Octane SDD EDS.
Porosity analysis of the cold sprayed coating was performed both on
Mechanical properties were investigated by Vickers microhardness as-sprayed and heat-treated samples. Porosity was evaluated on digital
measures, dynamic nanoindentation and four-points bending tests. images of cross sections (ASTM E-2109-01) [20] at 200× magnification
A subset of samples was exposed in furnace in air at 300, 500, 650 using a Nikon Eclipse L150 optical microscope (Nikon Instruments B.V.,
and 750 °C for 30 min, and then cooled in air. The heating ramps and Amsterdam, The Netherlands) assisted by the LUCIA Measurement
dwell times simulated the heating sequence and temperature exposures image analysis software: the captured images were digitally binarized
during the following four-point bending tests. These test temperatures by applying a grey-scale threshold and the porosity amount (black
were specifically selected to investigate the mechanical behaviour of pixels) was evaluated by the software. Only the porosity evaluated in
the alloy, both below and above the fcc-to-hcp martensitic transfor- CGS samples was relevant to the present investigation, as the micro-
mation occurring at around 410 °C [9]. The temperature of 650 °C was structure of the PTA-deposited Stellite 6 was characterised by a very
selected as an intermediate temperature to investigate the behaviour of low porosity (< 0.1%).
the coating between the post-transition temperature (500 °C) and the Tribological behaviour of both CGS and PTA coatings was in-
temperature around which Stellite alloys still maintain their micro- vestigated by wear test and SEM analysis on wear tracks.
hardness, 750 °C [10]. Wear tests were performed with a TE53 SLIM/8941 multi-purpose
Heating steps and dwell times are reported in Table 4. tribometer (Phoenix Tribology Ltd., Newbury, UK), following the ASTM
Young's modulus was evaluated by dynamic indentation (ISO G77 standard test method for friction and wear testing in inverted
14577: 2007) [11] using a Nanotest instrumented indentation system block-on-ring configuration [21]. A stationary cubic-shaped block
(MicroMaterials Ltd., Wrexham, UK) [12] both on as-sprayed and heat- (12.5 mm side), previously ultrasonically cleaned in acetone and
treated samples. Two sets of loads (500 and 1000 mN) were applied for weighted, was pressed against a rotating, alumina-coated 60 mm steel
10 s, using a Berkovitch 3-sided indenter. Both loading and unloading ring. The test was performed at room temperature by applying dry
velocity was fixed at 50 mN/s, while post-test analysis of load/pene- friction. Block temperatures were measured during wear tests using a k-
tration curves was carried out using the approach described by Oliver type thermocouple. Two loads and two distances were selected with a
and Pharr [13]. The reduced elastic modulus is calculated by the constant slide velocity of 318 rpm (~1 m/s), resulting in a total of four
method shown in Eq. (1): sets of test parameters (Table 3). Three samples per set of parameters
dp were tested.
1
Er = , Wear tests were followed by weight and volume variation calcula-
2 A dh (1)
tions. The specimens were cleaned and weighted, then worn out vo-
where A is the indenter area, and dp/dh is the elastic contact stiffness lumes were evaluated through a profilometric analysis using Talyscan

323
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Table 4
Wear test parameters.
Load [N] 150 300

Distance [km] 1 2
Slide velocity [rpm] 318

150 profilometer (Taylor Hobson, Leicester, UK) equipped with a laser


probe. Both weight and volume losses were calculated considering a
5 mm wide central zone for each wear track, to avoid geometrical and
experimental errors due to small misalignments, different sample di-
mensions, relative position of the block over the ring, etc.
EDS analysis were performed on wear tracks from both type of
deposited Stellite 6, to investigate whether a material transfer from the
alumina ring to the block occurred or not.

4. Results and discussion

4.1. Dynamic indentations

4.1.1. Young's modulus


Young's moduli of both as-deposited and heat treated self-standing
samples were measured at room temperature by dynamic indentation.
Results are reported in Fig. 2. E-moduli values for both CGS and
PTA-deposited Stellite 6, calculated at 500 and 1000 mN, are very

Fig. 1. High temperature four-point bending test machine: (a) general view; (b)
detail of the SiC flexure device.

Table 3
Heating steps and dwell times for heat treatments on cold-sprayed and PTA-
deposited samples. Tf is the final temperature.
Heating ramp step 1 From RT up to 40 °C at 5 °C/min

2 From 40 °C up to Tf - 50 °C at
15 °C/min
3 From Tf - 50 °C up to Tf - 5 °C at
10 °C/min
4 From Tf - 5 °C up to Tf at 5 °C/min

High temperature dwell time [min] 300 °C 15


500 °C 30
650 °C 40
750 °C 45
Fig. 2. Comparison of Young's moduli for cold sprayed and PTA-deposited
Stellite 6 evaluated by dynamic indentation calculated at (a) 500 mN, and (b)
1000 mN.

324
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Fig. 4. Comparison of Vickers microhardness for cold sprayed and PTA-de-


posited samples measured with normal loads of 50 gf.

Fig. 3. Comparison of Vickers microhardness for cold sprayed and PTA-de-


posited Stellite 6 evaluated by dynamic indentation calculated at (a) 500 mN,
and (b) 1000 mN.

similar for both loads; samples heat treated at 650 °C exhibit the higher
elastic moduli, especially the PTA-deposited coatings. As-sprayed cold-
sprayed Stellite 6 reaches its minimum E-modulus value.
Measurements of the elastic moduli calculated by dynamic in-
dentation for both CGS and PTA-deposited Stellite 6 involve very low
volumes of material, and do not consider effects related to possible
defects and/or porosity. This means that the mechanical properties
provided by this method are expressed on a microscopic scale. On the
other hand, four-point bending tests not only involve higher volumes of
material with respect to dynamic indentation, but the resulting elastic
moduli consider the combination of plastic deformation and the effect
of the aforementioned defects: such values express a mechanical
property on a macroscopic scale.

4.1.2. Microhardness
Microhardness values were obtained from dynamic indentation for
both type of coatings, at 500 and 1000 mN. Results are shown in Fig. 3.
Cold sprayed Stellite 6 exhibits higher value of microhardness Fig. 5. Stress-strain curves for (a) cold-sprayed, and (b) PTA-deposited Stellite
comparing to the values for PTA samples. Recalling the microscopic 6.

325
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

nature of such property, it makes sense assuming that the obtained dilution zone, to investigate whether the microhardness value is influenced
values represent a measure of the densification of the material fol- or not by such zone. Results are shown in Fig. 4.
lowing the heat treatment. As expected, cold sprayed Stellite 6 reaches high values of Vickers
Lower values of microhardness are registered for PTA coating over microhardness. Cold spray deposition, in fact, implies very high plastic
the range of heat treatment temperature. deformation of the particles on impact, providing the resulting coatings
with an elevated dislocation number and consequent hardening.
4.2. Vickers microhardness The overall trend matches the one obtained from dynamic in-
dentations.
Evaluation of Vickers microhardness was performed on cross sections of As for the PTA samples, values of microhardness measured in the
samples at room temperature after the heat treatments. For PTA-deposited coating are 40–50% lower than the ones measured in the cold sprayed
coatings, indentations were performed both on the coating zone and on the material, and they drop further in the dilution zone.
Such results are consistent with experimental data reported in other
studies [9,22].

4.3. Four-point bending test

Results of the mechanical characterization through four-point


bending test are reported in Fig. 5, depicting the stress-strain curves
obtained for the selected test temperatures.
Cold-sprayed Stellite (Fig. 5a) exhibits a brittle behaviour up to
500 °C; there is little to no difference between curves ranging from RT
to 500 °C. At 650 °C the stress-strain curve shows a plastic trend before
breaking at around 475 MPa (~0.47% strain), while the alloy plasti-
cally deforms at 750 °C. This behaviour can be explained with the cold
spray deposition process, by which a coating rich in interlamellar
cracks and deformed metal grains, and with little residual ductility is
obtained. These features lead to a crack propagation that results in a
brittle behaviour over a wide range of temperature.
PTA-deposited Stellite exhibits a plastic behaviour for all the test
Fig. 6. Comparison of elastic moduli of cold sprayed and PTA-deposited Stellite temperatures, as proved by its stress-strain curves in Fig. 5b. This can be
6 evaluated by four-point bending test at different test temperatures. easily ascribed to the PTA manufacturing process, that provides the
alloy with a microstructure that is very similar to the as-cast one.

Fig. 7. EDS map of post-etching cold-sprayed Stellite 6 in as-sprayed condition, 1000 X.

326
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Fig. 8. Post-etching microstructure of cold-sprayed Stellite 6 both (a) as-sprayed, and after heat treatment at (b) 300 °C, (c) 500 °C, (d) 650 °C, and (e) 750 °C, 500 X.

Fig. 6 shows the comparison of the Young's moduli for both CGS and deformation from 650 °C. [23] The low value of E-modulus reached at
PTA-deposited coatings. Elastic moduli for the cold-sprayed material room temperature by the CGS samples may be given by the only con-
reach lower values for all the test temperatures than the moduli of PTA tribution of voids in the microstructure. On the other hand, densifica-
samples. Modulus in cold sprayed coatings tends to slightly increase tion does not occur in PTA-deposited Stellite 6 as its structure is already
with temperature up to 500 °C, then its value lightly decreases at 650 °C very dense, so only the contribution from the elastic phenomena could
to finally drop at 750 °C, when the specimen is plastically deformed. be accountable for the decreasing trend of the elastic modulus with the
PTA-obtained Stellite 6 behaves differently over temperature, as temperature.
Young's modulus continuously decreases over the test temperature.
Such a difference in elastic modulus values and trends may be due 4.4. Microstructural and porosity evaluation
to a combination of two different effects acting simultaneously on cold
spray samples: i) the progressive densification – given both by the de- The microstructure of the as-deposited and heat-treated coatings
creasing porosity and increasing of cohesion among the grains –, that was investigated by etching.
leads to an apparent increasing of the E values with temperature; ii) the SEM backscattered images of CGS Stellite 6 after the etching are
interlamellar sliding phenomena occurring with increasing tempera- depicted in Figs. 7 and 8.
ture, that lead to a decreasing elastic modulus and to plastic By the EDS map in Fig. 7 for the as-sprayed sample it is possible to

327
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Fig. 9. Post-etching microstructure of PTA-welded Stellite 6 both (a) as-deposited, and after heat treatment at (b) 300 °C, (c) 500 °C, (d) 650 °C, and (e) 750 °C, 500 X.

identify the distribution of cobalt, nickel (light grains) and chrome interlamellar voids resembling cracks, tends to slightly increase with
(dark zones) throughout the microstructure. the test temperature, with a final decreasing at 750 °C. Such trend could
As the temperature of the heat treatment increases, the grains tend be explained by the progressive densification of the cold-sprayed mi-
to progressively agglomerate forming a denser structure (Fig. 8a–e). crostructure, that tends to open voids as the grains agglomerate
Fig. 9 shows the effect of etching on PTA-deposited Stellite 6. The (Fig. 8b–c), to reach a final density that is very similar to the initial one.
coating exhibits its peculiar dendritic structure, which is maintained
throughout the heat treatment. The cobalt grains tend to slightly gather
in a coarse structure only at higher temperatures (Fig. 9e), but such 4.5. Tribology
evolution is not as evident as in the cold-sprayed alloy.
By comparing the microstructures showed in Figs. 8 and 9, there is a Results from the tribological tests were elaborated in terms of fric-
considerable difference in the density of both structures. A measure of tion coefficient, block temperature, mass loss, and volume loss.
porosity is given in Fig. 10. The friction coefficients obtained for each test condition and ma-
Porosity in cold sprayed samples, part of which is given by terial are shown in Fig. 11.
The friction coefficients for both cold sprayed and PTA-deposited

328
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Fig. 10. Porosity evaluation of cold-sprayed Stellite 6 on as-sprayed and heat-


treated samples.

Fig. 12. Comparison of block temperatures gathered from tribological tests


performed for (a) 2 km and (b) 1 km sliding distance.

values both at 300 and at 150 N load (~0.40 ÷ 0.45, dark blue and
light blue lines in Fig. 11a, b).
PTA-deposited coatings exhibit a different behaviour with respect to
the applied load: while, in fact, a ~0.40 ÷ 0.45 stable value is reached
at 300 N for both 2 and 1 km distance (red lines in Fig. 11a, b) – si-
milarly to the results gathered from the cold sprayed samples -, friction
coefficient drops to ~0.20 ÷ 0.25 at 150 N load.
Such results are correlated with the ones obtained from block
temperature measurements, as shown in Fig. 12: while, in fact, cold
sprayed Stellite exhibits higher friction coefficients if compared to PTA
samples, the relative block temperature increases accordingly, because
of the more intense dissipation phenomena due to the higher friction
force. It is worth noting that, for both type of samples and test distance,
the higher the applied loads, the higher the block temperature.
Weight and volume variation are depicted in Fig. 13.
Both cold sprayed and PTA-deposited Stellite 6 coatings exhibit the
Fig. 11. Comparison of friction coefficients gathered from tribological tests same trend of weight variation depending on test distance (Fig. 13a): as
performed at the constant velocity of 1 m/s for (a) 2 km and (b) 1 km sliding expected, the shorter the running distance, the lesser the lost mass.
distance. Anyway, PTA-deposited coatings are affected by a mass loss that is
higher than CGS ones, and this difference is more evident in samples
Stellite 6 are characterised by a rapid transient state through the time tested for 2 km running distance. The relatively low hardness of PTA
for each test condition, in terms of both distance and applied load. Stellite 6, compared to the one of cold-sprayed coating, could be ac-
Specifically, coefficients of CGS coatings reach very similar stable countable for the evident mass variation; moreover, the increase in

329
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

hand, the PTA-deposited alloy (Fig. 14b) exhibits a massive presence of


Al in its wear track, suggesting that a material transfer occurred from
the alumina ring to the block. Moreover, an evidence of the exposure of
the dilution zone is given as the Fe peak corresponding to point 2 in-
creases.

5. Conclusions

A comparative study of the tribological behaviour and high tem-


perature mechanical properties of Stellite 6 obtained by cold spray and
PTA welding was carried out.
The interest in the cold spray technique is given by the chance to
obtain a coating that is oxide-free and lacks the dilution zone, that is a
typical feature of the PTA welding and may affect its tribological per-
formance.
The investigation, while preliminary, highlights some important
differences between the two type of coatings.
Namely, the cold-sprayed alloy exhibited high microhardness in
both micro- and nano-indentation measurements, which actually re-
flected in satisfactory tribological performances. Very low consumption
rates – expressed in terms of weight variation and volume loss – were
registered for CGS samples in all test conditions, while PTA specimens
were affected by important material loss, especially at 300 N load.
Moreover, alumina transfer occurred from the ring to the PTA-welded
block, suggesting that this type of coating suffers an intense adhesive
wear probably due to its high ductility at the test temperatures, that
leads to the inclusion of debris from the rotating counterpart into the
wear track.
Mechanical properties of the cold-sprayed Stellite 6 both at room
and at high temperature were, on the other hand, less impressive when
compared with the ones exhibited by the corresponding PTA coating.
Results from the four-point bending tests showed that the CGS samples
experienced a brittle behaviour up to 500 °C, probably to be attributed
to the deposition process that gives a coating rich in interlamellar crack
and plastically deformed particles. Such features may lead to a crack
Fig. 13. Comparison of material consumption after tribological tests expressed propagation given by a progressive structure densification that results
in terms of (a) weight, and (b) volume variations. in the brittle curves represented in Fig. 6a. At 650 °C and especially at
750 °C, plastic deformation mechanisms activate and overcome the ef-
fects of the densification.
mass loss rate experienced by PTA samples tested for 2 km may be
On the other hand, PTA-deposited Stellite 6 behaves as an as-cast
explained by the progressive material consumption that exposes the soft
alloy due to the manufacturing process. This aspect results in a plastic
dilution zone during the wear test.
behaviour in every test condition, to room temperature up to 750 °C.
While there is a very limited difference in weight variation between
Elastic moduli extracted from stress-strain curves for both kind of
samples tested for 1 km of sliding distance for both type of coatings
coatings reflect the result, while different observations must be con-
(around 0.02 g between 300 and 150 N of applied load for both CGS and
sidered for the elastic moduli calculated by dynamic indentations.
PTA-deposited materials), after 2 km PTA blocks experience a more
Recalling that the latter provide properties evaluated on a microscopic
evident higher mass loss if compared to the CGS ones at the same test
scale, values of Young's moduli calculated for the cold-sprayed samples
conditions (around 0.05 g difference between 300 N and 150 N applied
are higher than the ones collected for the PTA-deposited material. This
load for PTA samples, versus 0.01 g difference registered for cold spray
fact may be due to the local densification of the cold-sprayed Stellite,
material). A suitable explanation for this difference in mass variations
which also leads to high values of Vickers microhardness that match the
could be given by the material consumption during the wear test that,
ones obtained by microindentation tests.
in the PTA-deposited alloy, progressively exposes the underlying softer
Evidence of the progressive densification experienced by the cold-
dilution zone, with a consequent increase of the mass loss rate.
sprayed material is provided by the post-etching microstructural ana-
The evaluation of worn out volumes (Fig. 13b) matches the results
lysis carried out on as-sprayed and heat-treated samples, Fig. 12. Be-
gathered from weight variation analysis, with PTA-deposited Stellite
cause of such a visible microstructural evolution throughout the treat-
experiencing major volume loss in every test condition.
ments at high temperature, and unlike the PTA-deposited coating, a
EDS analysis on wear tracks was performed, as shown in Fig. 14.
porosity evaluation was possible.
Punctual analysis on the wear track of the cold-sprayed Stellite 6
As a final remark, the results collected so far provide only a partial
(Fig. 14a), performed on three different points of the image, shows that
overview of the main differences that qualify cold-sprayed and PTA-
there is a very limited material transfer from the alumina ring to the
deposited Stellite 6, as this investigation does not consider the possible
block, as the Al content at each analyzed point is very low. On the other
effect on mechanical properties and tribological behaviour of the stress-

330
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

Fig. 14. EDS punctual analysis on wear tracks: (a) cold-sprayed, and (b) PTA-deposited Stellite 6.

induced fcc-to-hcp transition that concerns the whole family of cobalt- alloy for both type of deposition techniques. Moreover, the evaluation
base alloys. It would be of special interest investigating whether the of the effects of the deposition parameters and the consequent hard-
combination of high temperature and stress resulting from four-point ening of the CGS coatings on the transition temperature would be an-
bending tests would result in a phase transition, and in what terms such other interesting subject for the following experimentations.
a transformation would affect the overall mechanical behaviour of the

331
L. Baiamonte, et al. Surface & Coatings Technology 371 (2019) 322–332

References indentation: advances in understanding and refinements to methodology, J. Mater. Res.


19 (1) (2004) 3–20.
[14] ASTM B596-89(2017), Standard Specification for Gold-Copper Alloy Electrical Contact
[1] A. Motallebzadeh, E. Atar, H. Cimenoglu, Sliding wear characteristics of molybdenum Material, ASTM International, West Conshohocken, PA, 2017, www.astm.org.
containing Stellite 12 coating at elevated temperatures, Tribology Int 91 (2015) 40–47. DOI:https://doi.org/10.1520/B0596-89R17.
[2] Y.J. Kim, Wear and corrosion resistance of PTA weld surfaced Ni and Co based alloy [15] ASTM C1161-18, Standard Test Method for Flexural Strength of Advanced Ceramics at
layers, Surf. Eng. 15 (1999) 495–501. Ambient Temperature, ASTM International, West Conshohocken, PA, 2018, www.astm.
[3] Š. Houdková, E. Smazalová, Z. Pala, Effect of heat treatment on the microstructure and org. DOI:https://doi.org/10.1520/C1161-18.
properties of HVOF-sprayed Co-Cr-W coating, J. Therm. Spray Technol. 25 (2016) [16] ASTM C1211-18, Standard Test Method for Flexural Strength of Advanced Ceramics at
546–557, https://doi.org/10.1007/s11666-015-0365-5. Elevated Temperatures, ASTM International, West Conshohocken, PA, 2018, www.astm.
[4] R. Liu, Q. Yang, F. Gao, Tribological behavior of Stellite 720 coating under block-on-ring org. DOI:https://doi.org/10.1520/C1211-18.
wear test, Mater. Sci. Appl. 3 (2012) 756–762, https://doi.org/10.4236/msa.2012. [17] G. Di Girolamo, F. Marra, C. Blasi, M. Schioppa, G. Pulci, E. Serra, T. Valente, High-
311110. temperature mechanical behavior of plasma sprayed lanthanum zirconate coatings,
[5] M. Brandt, S. Sun, N. Alam, P. Bendeich, A. Bishop, Laser cladding repair of turbine Ceram. Int. 40 (2014) 11433–11436, https://doi.org/10.1016/j.ceramint.2014.03.110.
blades in power plants: from research to commercialisation, Int. Heat Treat. Surf. Eng. [18] G. Pulci, M. Tului, J. Tirillò, F. Marra, S. Lionetti, T. Valente, High temperature me-
(2009), https://doi.org/10.1179/174951409X12542264513843. chanical behavior of UHTC coatings for thermal protection of re-entry vehicles, J. Therm.
[6] Y.P. Kathuria, Some aspects of laser surface cladding in the turbine industry, Surf. Spray Technol. 20 (2011) 139–144, https://doi.org/10.1007/s11666-010-9578-9.
Coatings Technol. 132 (2000) 262–269, https://doi.org/10.1016/S0257-8972(00) [19] R. Lupoi, A. Cockburn, C. Bryan, M. Sparkes, F. Luo, W. O'Neill, Hardfacing steel with
00735-0. nanostructured coatings of Stellite-6 by supersonic laser deposition, Light Sci. Appl.
[7] B.-H. Yoon, C.-H. Lee, H.-J. Kim, Effect of dilution on wear performance of plasma (2012), https://doi.org/10.1038/lsa.2012.10.
transferred arc deposited layers, ISIJ Int. 57 (2017) 913–920. [20] ASTM E2109-01, Test Methods for Determining Area Percentage Porosity in Thermal
[8] N. Cinca, J.M. Guilemany, Structural and properties characterization of stellite coatings Sprayed Coatings, ASTM International, West Conshohocken, PA, 2001, www.astm.org.
obtained by cold gas spraying, Surf. Coatings Technol. 220 (2013) 90–97, https://doi. DOI:https://doi.org/10.1520/E2109-01.
org/10.1016/j.surfcoat.2012.11.026. [21] ASTM G77-17, Standard Test Method for Ranking Resistance of Materials to Sliding Wear
[9] N. Cinca, J.M. Guilemany, Cold gas sprayed Stellite-6 coatings and their wear resistance, Using Block-on-Ring Wear Test, ASTM International, West Conshohocken, PA, 2017,
J. Mater. Sci. Eng. 02 (2013), https://doi.org/10.4172/2169-0022.1000122. www.astm.org. DOI:https://doi.org/10.1520/G0077-17.
[10] S. Kapoor, High-Temperature Hardness and Wear Resistance of Stellite Alloys, (2012) [22] Q.Y. Hou, J.S. Gao, F. Zhou, Microstructure and wear characteristics of cobalt-based alloy
(ISBN: 978-0-494-91596-7). deposited by plasma transferred arc weld surfacing, Surf. Coatings Technol. 194 (2005)
[11] ISO 14577, Instrumented indentation test for hardness and materials parameters, S.l.: 238–243, https://doi.org/10.1016/j.surfcoat.2004.07.065.
ISO, (2002). [23] L. Baiamonte, F. Marra, G. Pulci, J. Tirillò, F. Sarasini, C. Bartuli, T. Valente, High tem-
[12] P. Palmero, G. Pulci, F. Marra, T. Valente, L. Montanaro, Al2O3/ZrO2/Y3Al5O12 compo- perature mechanical characterization of plasma-sprayed zirconia-yttria from conven-
sites: a high-temperature mechanical characterization, Materials (Basel) 8 (2015) tional and nanostructured powders, Surf. Coatings Technol. 277 (2015) 289–298, https://
611–624, https://doi.org/10.3390/ma8020611. doi.org/10.1016/j.surfcoat.2015.07.071.
[13] W.C. Oliver, G.M. Pharr, Measurement of hardness and elastic modulus by instrumented

332

You might also like