Sie sind auf Seite 1von 16

Surface & Coatings Technology 206 (2012) 4079–4094

Contents lists available at SciVerse ScienceDirect

Surface & Coatings Technology


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

Cermet coatings with Fe-based matrix as alternative to WC–CoCr: Mechanical and


tribological behaviours
Giovanni Bolelli a,⁎, Tim Börner a, Francesco Bozza a, Valeria Cannillo a, Gennaro Cirillo b, Luca Lusvarghi a
a
Department of Materials and Environmental Engineering, University of Modena and Reggio Emilia, Via Vignolese 905, I-41125 Modena (MO), Italy
b
Parma Spray Italia S.r.l., Via Giovanni XXIII, I-43040 Varano de'Melegari (PR), Italy

a r t i c l e i n f o a b s t r a c t

Article history: Recently, cermet coatings with Fe-based metal matrix have emerged as a less hazardous and more environ-
Received 20 February 2012 mentally friendly alternative to WC–Co-based ones, which have known inhalation toxicity problems. This
Accepted in revised form 31 March 2012 study therefore aimed to validate WC-based cermet coatings with Fe-based matrix, obtained using a com-
Available online 10 April 2012
mercially available feedstock powder, as an alternative to WC–CoCr.
HVOF-sprayed WC–15 wt.%FeCrAl layers were therefore obtained using different oxygen and fuel (kerosene)
Keywords:
WC–FeCrAl cermet
flow rates and powder feed rates; their mechanical and tribological properties were compared to HVOF-
High Velocity Oxygen-Fuel (HVOF) spraying sprayed WC–10 wt.% Co–4 wt.%Cr.
Sliding wear The WC–FeCrAl coatings always exhibited equi-biaxial compressive residual stress state and possessed dense
Abrasive wear microstructures, with homogeneous metal matrix, but they contained more oxide inclusions than WC–CoCr.
Residual stress Their characteristics were significantly affected by the normalised oxygen-fuel ratio (λ).
Cyclic impact behaviour Small but meaningful differences existed between the ball-on-disc sliding wear rates of the various WC–
FeCrAl coatings, the best sample being that with the most favourable combination of compressive residual
stress, low oxidation and high hardness/modulus (H/E) ratio. Its sliding wear resistance was comparable to
that of WC–CoCr. The cyclic ball impact resistance of WC–FeCrAl layers was also comparable to that of
WC–CoCr, but the dry particle abrasion resistance was inferior, because of the brittleness induced by the
oxide inclusions.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction for inclusion in the list of suspect human carcinogens as well [14]. On
the other hand, Ni- and Co-containing coatings sometimes fail to con-
Cermets consisting of WC particles embedded in a Co- or Ni-based form to the requirements of some very specific but economically
metal matrix or of CrxCy in a Ni-based matrix are frequently employed significant application fields, most notably the food and beverage pro-
as coating materials for the protection against various forms of wear cessing industry, where usable materials must meet stringent
(sliding, abrasive, erosive, etc.) in industrial machinery [1–8]. Ther- requirements. Specifically, strict limitations exist against the potential
mal spray processes, especially those involving high kinetic energy, contamination of the products with toxic or harmful elements,
such as the High Velocity Oxygen-Fuel (HVOF) or High Velocity Air- released in the form of fine wear debris particles or of ions, leached
Fuel (HVAF) techniques, are usually the preferred deposition in liquid media [15,16].
methods, on account of the rather high productivity and of the high Searching for replacements to Ni- and Co-based coating materials,
quality of the resulting layers, which possess low porosity, high cohe- capable of overcoming the above-mentioned shortcomings, the au-
sive strength and satisfactory adhesion to the substrate [9,10]. thors recently investigated the properties of HVOF-sprayed Fe-based
These coatings, however, pose some safety and health issues. First metal alloy coatings [17]. Although they were shown to be viable al-
of all, handling of the feedstock powder in thermal spray shops im- ternatives for electroplated hard chromium and for other thermally-
plies some hazards. Ni-based alloys are allergenic and are labelled sprayed metal alloy coatings, their wear resistance could not compare
as suspect carcinogenic agents, as seen e.g. in [11]: for instance, Ni- with that of cermet materials. These coatings, therefore, would not be
based powders come under the H351 hazard statement according to suitable in applications where the tribological characteristics of con-
European Commission regulation EC 790/2009 [12]. WC–Co materials ventional cermets, such as WC–Co-based ones, would be required.
are also toxic by inhalation [13] and are currently under consideration A better solution for this purpose might be offered by thermal
spray-grade cermet powders free of Ni and Co, consisting of WC par-
ticles in a Fe alloy matrix, which have recently been introduced in the
⁎ Corresponding author. Tel.: + 39 0592056233; fax: + 39 0592056243. market [11]. The properties of the thermally-sprayed coatings
E-mail address: giovanni.bolelli@unimore.it (G. Bolelli). obtained from those powders were considered in a few publications

0257-8972/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2012.03.094
4080 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

[13,18], but a complete and systematic investigation has not been Table 2
performed yet. HVOF process parameters.

The purpose of the present paper, therefore, was to provide a com- Parameter O2 flow rate Kerosene flow λa Powder feed Stand-off
prehensive mechanical and tribological characterisation of WC– set (SL/min) rate (L/h) rate (g/min) distance (mm)
FeCrAl coatings deposited by liquid-fuelled HVOF-spraying technique. #1 950 23 1.27 100 400
Their microstructural features, depth-sensing indentation response, #2 850 18 1.46 100
residual stresses, sliding and abrasive wear behaviour and cyclic im- #3 950 18 1.63 60
#4 850 23 1.14 60
pact resistance were investigated as a function of some key spraying
#5 900 21 1.32 80
parameters and were compared to those of a standard WC–CoCr
a
layer, employed as a term of comparison. λ = normalised oxygen fuel ratio, computed based on a simplified kerosene com-
bustion reaction C12H26(l) + 37/2 O2(g) → 12 CO2(g) + 13 H2O(g) with ρ = 0.8 g/cm3
as density of kerosene and ρ = 1.4290 kg/m3 as standard density of O2.
2. Materials and methods
size analyser with Hydro-S wet dispersion unit, Malvern Instruments
2.1. Coating deposition
Ltd., Malvern, UK).
Commercially available thermal spray-grade WC–15 wt.% FeCrAl
2.3. Characterisation of coatings
and WC–10 wt.% Co–4 wt.% Cr feedstock powders, manufactured by
agglomeration and sintering, were employed (Table 1). The sub-
Cross-sections of the WC–FeCrAl and WC–CoCr coatings were pre-
strates were C40 plain carbon steel plates (100 × 100 × 5 mm) and
pared by metallographic cutting, hot-mounting in phenol resin,
AA 6082T6 aluminium alloy plates (80 × 80 × 8 mm); the substrates
grinding with diamond papers (up to 5 μm particle size) and polish-
were preliminarily degreased with acetone, grit-blasted using
ing using diamond slurries (up to 3 μm particle size). Polished
240 μm alumina grits and pre-heated at 60 °C for 1 h in a drying
cross-sections were employed for microstructural observations
oven to remove adsorbed moisture.
through SEM and for qualitative and semi-quantitative chemical anal-
The powers were sprayed onto the substrate plates using a
ysis through EDX. Specifically, semi-quantitative assessments of
kerosene-fuelled Praxair-Tafa JP5000 HVOF torch, equipped with a
chemical compositions were obtained by performing large-area EDX
152.4 mm-long nozzle. As specified in Table 1, the WC–FeCrAl pow-
scans on two cross-sectional views at 400× magnification for each
der was deposited onto the carbon steel plates using 5 distinct pa-
coating, with electron beam energy of 15 keV.
rameter sets (labelled from #1 to #5), which differed in oxygen and
The amount of oxide inclusions in each coating was assessed by
kerosene flow rates and powder feed rates (Table 2), in order to
image analysis technique (software: ImageJ version 1.43u, NIH, Be-
study their influence on coating properties. An additional WC–FeCrAl
thesda, USA) onto 5 backscattered electron (BSE)-SEM micrographs
coating was deposited onto the aluminium alloy plate using parame-
at 1000× magnification. Coating thickness was also assessed by
ter set #1, in order to check possible influences of substrate proper-
image analysis on low magnification micrographs.
ties on the characteristics of the coating.
Depth-sensing Berkovich micro-indentation tests were also per-
Parameter set #1, representing standard HVOF processing condi-
formed onto the polished cross-sections, with maximum load of 3 N,
tions for WC–CoCr cermets, was also employed in order to spray the
loading/unloading rate of 2.4 N/min, holding time at maximum load
WC–CoCr feedstock powder onto the carbon steel plates, in order to
of 15 s (Micro-Combi Tester, CSM Instruments, Peseux, Switzerland).
obtain reference coatings, which serve as a term of comparison to
Instrumented Vickers hardness and elastic modulus were assessed
evaluate the characteristics of the WC–FeCrAl layers.
using the Oliver–Pharr procedure [19], in accordance with the ISO
14577 standards.
2.2. Characterisation of the WC–FeCrAl powder Square samples of 23 × 23 mm were cut from the coated plates
and polished using diamond papers and diamond slurries as de-
The phase composition of the WC–FeCrAl powder was analysed by scribed above, achieving a final roughness Ra ≈ 0.02 μm (assessed
X-ray diffractometry (XRD: X'Pert PRO diffractometer, PANAlytical, through optical confocal profilometry: Conscan profilometer, CSM In-
Almelo, The Netherlands), using Cu–Kα radiation emitted from an struments). The thickness reduction during grinding and polishing
X-ray tube operated at 40 kV energy and 40 mA current, and employ- was assessed using a digital outside micrometre.
ing a gas-proportional X-ray detector equipped with single-crystal Phase composition of the coatings was assessed onto polished sur-
monochromator. A 2θ range from 20° to 100° was scanned with a faces by XRD, using the same experimental set-up described in
step size of 0.025° and an acquisition time of 0.90 s/step. Section 2.1.
Cross-sections of the powder particles, obtained by cold-mounting Residual stress was also measured on the polished surfaces by X-
in polyester resin followed by grinding using SiC papers (up to 2500 ray diffractometry. The classical sin 2ψ technique was employed
mesh) and polishing with diamond slurries (up to 3 μm size), were [20]: 11 different ψ-tilt values symmetrically distributed in the
observed by scanning electron microscopy (SEM: XL-30, FEI, Eindho- [−45°; + 45°] range, corresponding to sin 2ψ values of 0, 0.1, … ,
ven, The Netherlands); qualitative and semi-quantitative chemical 0.5, were adopted in order to follow the displacement of the (112)
analyses were also obtained by energy-dispersive X-ray microanaly- peak of WC, located at 2θ ≈ 98.7° (according to JCPDF 51-939).
sis (EDX: INCA, Oxford Instruments Analytical, Abingdon, UK). Using the same diffractometer mentioned above in a point-focus con-
The particle size distribution was assessed by laser diffraction figuration, a 2θ scan from 97.0° to 100.5°, with a step size of 0.017°
technique using wet dispersion method (Mastersizer 2000 particle and an acquisition time of 151.5 s/step, was performed at each ψ-tilt
position. The complete ψ-scan was repeated along three different φ
Table 1 orientations (0°, 45°, 90°).
Characteristics of feedstock powders (manufacturer: H.C. Starck, Laufenburg, The normal stress along each φ orientation was computed using
Germany). the equation for plane-stress condition (1) [20]:
Composition (wt.%) WC–15(Fe–20Cr–7Al) WC–10Co–4Cr
Powder trade name Amperit 618.074 Amperit 556.074
Particle size range (μm) − 45 + 15 − 45 + 15 dφψ −d0 ðhklÞ 1 ðhklÞ 2
ε φψ ¼ ¼ S1 ðσ 1 þ σ 2 Þ− S2 σ φ sin ψ: ð1Þ
Process parameter sets (see Table 2) #1–#5 #1 d0 2
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4081

Where: The volume wear loss of the sample was assessed by optical
confocal profilometry (Conscan profilometer, CSM Instruments)
dφψ interplanar spacing of the selected diffraction peak ((112) and was converted to wear rate, i.e. volume loss per unit nor-
peak of WC), measured at a given φ,ψ position; mal load and unit sliding distance, according to Archard equa-
d0 interplanar spacing of the selected diffraction peak ((112) tion [23]. The worn surface of the sample was also inspected
peak of WC) in the unstressed material; by SEM.
σφ normal stress along the φ direction; Dry particle abrasion tests were performed on as-deposited
ðhklÞ  ðhklÞ 
S1 ¼ Ev ðhklÞ ; 12 S2 ¼ 1þv
E ðhklÞ X-ray elastic constants along the coating surfaces (i.e. without polishing) using a dry sand-steel
(hkl) direction. wheel apparatus (a modified version of the ASTM B611-85(2005)
test method for abrasive wear resistance of cemented carbides),
Similar to the measurement procedure described in [21], the previously employed by the authors in [24,25] as well. The sample
coating's elastic modulus E assessed by depth-sensing indentation was pressed with a 40.2 N load against a 200 mm-diameter Fe360
was employed in Eq. (1), while the Poisson's ratio was assumed steel wheel rotating at 75 rpm, with a tangential flow (75 g/min)
to be υ = 0.3: although this may imply a certain degree of approx- of angular alumina particles (180 μm average size). A total of 90
imation, the results are certainly useful in order to rank the residual disc revolutions were performed (corresponding to a wear distance
stress magnitude in the various coated samples. of 56.5 m). All samples were tested at least three times. The vol-
The unstressed interplanar spacing d0 was taken from the XRD ume wear loss of the samples was computed based on the mea-
pattern of the WC–FeCrAl feedstock powder. In any case, the error sured wear scar length and was converted to wear rate as
on σφ due to uncertainties on d0 is generally of ≈0.1% [22]. Accord- described previously. The wear scars were also observed by SEM
ingly, it was ascertained that, if this d0 value was replaced with the (Quanta 200, FEI).
dψ = 0° value in the analysis of the present results, the change in the Cyclic ball drop impact tests were performed on the polished
value of σφ was practically negligible. 23 × 23 mm square samples: a 39 mm-diameter X200Cr11 steel
Dry sliding wear tests were performed on the polished, 23× 23 mm- ball, attached to an overall mass of 1.35 kg, was dropped 200
sized square samples under ball-on-disc configuration (according to times (3 Hz impact frequency) onto a fixed location on the sample
ASTM G99-05), using a pin-on-disc tribometer (High-Temperature surface. The schematic of the test equipment is shown in [26]. The
Tribometer, CSM Instruments) equipped with 6 mm-diameter sintered surface morphology of the tested coatings was inspected by optical mi-
α-Al2O3 spheres (nominal Vickers hardness = 1900 HVN). Tests were croscopy and the volume of the impact crater was assessed by optical
performed at a temperature of 25 ± 2 °C with a relative humidity of confocal profilometry (Conscan profilometer). Cross-sectional mor-
58± 2%. The normal load was 10 N, the linear relative sliding speed of phologies were further inspected by SEM (Quanta-200): the samples
the ball on the sample was 0.10 m/s, the radius of the wear track was were embedded in epoxy resin, dissected by metallographic cutting,
7 mm and the overall sliding distance was 5000 m. Each sample was ground and polished with diamond papers and diamond slurries as de-
tested at least twice to ensure repeatability of the results. scribed previously.

Fig. 1. Cross-sectional backscattered electron (BSE)-SEM micrographs of the WC–FeCrAl feedstock powder (A: overview; B: magnified view of one particle; C: detail of WC particles)
and EDX microanalyses (D) acquired on the regions marked as “1” and “2” in panel B. In panel B: label 1 = WC-rich area; label 2 = Cr-rich area without WC particles.
4082 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

± standard deviation of the two large-area EDX scans acquired on


each coating, as specified in Section 2.3), matches reasonably well
with the nominal one listed in Table 1; however, significant amounts
of oxygen seem to have been incorporated, particularly in the WC–
FeCrAl layers. It should be remarked that oxygen concentrations in
Table 3 are to be considered as indicative values only, since the reli-
ability of a quantitative assessment of oxygen by the EDX technique
is limited. Some approximation in the chemical analysis in Table 3
(where the sum of all weight percentages was normalised to 100) is
due to the inability of the EDX technique to quantify C.
The clear distinction between bright WC-rich areas and dark Cr-
rich areas, which characterised the powder particles, is no more pre-
sent in all of the WC–FeCrAl coatings (Figs. 4A–E and 5). By contrast,
the WC–FeCrAl coatings exhibit a homogeneous matrix (Fig. 5) simul-
taneously containing Fe, Cr, Al and, presumably, some dissolved W
(as shown by the EDX spectrum in Fig. 6A). Consistently, the X-ray
Fig. 2. XRD patterns of the WC–FeCrAl feedstock powder and of corresponding HVOF- diffraction peaks of α-Fe and η-M6C are not detectable in the XRD
sprayed coatings. Legend: = WC (JCPDF 51-939); = W2C (JCPDF 35-776); = W
patterns of the WC–FeCrAl coatings (Fig. 2). The coatings also contain
(JCPDF 4-806) or (W,Cr)2C [32]; α = Fe — b.c.c. (JCPDF 6-696); η = M6C (JCPDF 27-
1125). a significant amount of oxide inclusions (easily recognisable from
their dark contrast in BSE micrographs, Fig. 4A–E), consistent with
the EDX results in Table 3. These inclusions appear both as elongated
3. Results and discussion stringers at lamellar boundaries (Fig. 5, label 1) and as small spherical
intralamellar particles (Fig. 5, label 2). The former are somewhat
3.1. Microstructural characterisation of powders and coatings enriched in Al compared to the base alloy composition, but contain
significant amounts of W, Cr and Fe as well, as shown by the EDX
The narrow, monomodal particle size distribution of the WC– spectra in Fig. 6B; the latter, by contrast, are extremely rich in Al
FeCrAl powder, characterised by d0.1 = 19.0 μm, d0.5 = 30.7 μm, and (Fig. 6B): probably, Al2O3 is their main constituent.
d0.9 = 47.0 μm, matches remarkably well with the nominal values in- All of these microstructural features clearly indicate that, during
dicated in Table 1. spraying, irrespective of the processing conditions, the matrix
Its particles exhibit a rounded, porous morphology (Fig. 1A), typ- phases (α-Fe and η-M6C) of the original WC–FeCrAl powder parti-
ical of agglomerated and sintered powders, as seen e.g. in [27,28]. cles were completely melted. During flight, the molten phase expe-
Backscattered electron views highlight a clear separation (better rienced interdiffusion, homogenisation, and some degree of oxidation;
seen in the higher magnification micrograph of Fig. 1B) between dar- then, upon impact onto the substrate, it solidified into an amorphous
ker regions, rich in Cr but devoid of WC (EDX spectra in Fig. 1D), and and/or nanocrystalline structure, as a consequence of the high quench-
brighter regions, rich in WC particles embedded in a Fe–Al matrix. ing rate. This phenomenon is analogous to the amorphisation of the
Specifically, micrometric and sub-micrometric WC particles, having metal matrix in thermally-sprayed WC–Co-based cermets [29].
quite angular morphology, co-exist in these regions (Fig. 1C). In-flight interaction of the molten matrix with the O2 left in the
This separation is reflected by the XRD pattern of the powder HVOF gas jet after the combustion of kerosene (all of the process pa-
(Fig. 2), where, apart from the obvious diffraction peaks of WC, rameters indeed feature oxygen-rich mixtures with λ > 1, as shown in
peaks ascribable to α-Fe (b.c.c. structure) and to M6C (η-phase) are Table 2, since the kerosene-fuelled JP5000 torch can only operate
detectable: they presumably correspond to the Fe–Al matrix sur- with excess oxygen) and/or with surrounding air entrained in the
rounding the WC particles and to the dark Cr-rich regions, respective- jet by turbulent mixing causes the formation of oxides. The oxides
ly. This means the η-phase, in this case, consists of a (Cr,Fe)6C are originally developed on the surface of the droplet and are then
structure. dragged towards the centre because of turbulent flow within the mol-
After HVOF spraying, dense and homogeneous coatings, with ten metal, giving rise to the spherical, Al-rich inclusions seen inside
thicknesses lying between 130 μm and 220 μm (Table 3), free of any the splats in Fig. 5A and B. Their morphology is indeed identical to
major defect and of visible porosity, and macroscopically similar to that of oxide inclusions developed in flight inside thermally-sprayed
the WC–CoCr reference, are obtained (Fig. 3). By comparing the thick- metal droplets [30,31].
ness data in Table 3 and the process parameters in Table 2, it can be After impact quenching onto the substrate, the splats still remain
inferred that thickness is primarily affected by powder feed rate. at high temperature for a short time, and experience subsequent
The overall chemical composition of the coatings, determined by brief re-heating during successive passes of the torch. During these
semi-quantitative EDX analysis (Table 3: the values are the average periods, oxidation of the surface of the deposited lamella occurs,

Table 3
Chemical composition of the coatings from EDX analysis, index of carbide retention (I) computed from Eq. (2), oxide content and thickness from image analysis on BSE-SEM
micrographs.

Sample W (wt.%) Fe (wt.%) Cr (wt.%) Al (wt.%) O (wt.%) I Oxide content (vol.%) Thickness (μm)

#1 80.0 ± 0.6 10.8 ± 0.1 2.6 ± 0.4 0.6 ± 0.1 6.0 ± 0.4 0.78 7.3 ± 2.0 184 ± 6
#2 78.8 ± 0.5 11.7 ± 0.1 2.9 ± 0.2 0.8 ± 0.1 5.9 ± 0.4 0.89 9.6 ± 0.6 217 ± 8
#3 78.4 ± 0.1 11.0 ± 0.1 2.7 ± 0.1 0.9 ± 0.1 6.9 ± 0.1 0.83 7.9 ± 1.0 151 ± 9
#4 79.9 ± 0.3 10.9 ± 0.2 2.8 ± 0.1 0.7 ± 0.1 5.7 ± 0.3 0.75 6.1 ± 0.6 133 ± 10
#5 80.8 ± 0.9 10.9 ± 0.5 2.7 ± 0.3 0.6 ± 0.1 5.0 ± 0.1 0.81 5.5 ± 1.1 164 ± 8
WC–CoCr 81.7 ± 0.6 Co: 11.3 ± 0.2 4.4 ± 0.2 – 2.7 ± 0.1 0.88 1.8 ± 0.7 158 ± 12
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4083

Fig. 3. Cross-sectional BSE-SEM micrographs (overviews) of the HVOF-sprayed coatings deposited onto C40 steel plates: WC–FeCrAl #1 (A), #2 (B), #3 (C), #4 (D), #5 (E) and WC–
CoCr reference (F).

according to the mechanisms illustrated in [31], developing the splat- WC (Fig. 5A, label 3: the corresponding EDX spectrum in Fig. 6A con-
boundary oxide stringers also seen in Fig. 5A and B. firms that these areas are richer in W than ordinary matrix areas).
This interpretation of the oxidation mechanisms, based on exten- Further evidence comes from the XRD patterns of the coatings,
sive previous literature studies [30,31], accounts for the clear differ- which exhibit diffraction peaks of W2C witnessing some degree of
ence between the morphology and chemical composition of the two decarburisation of WC. A weak secondary peak at 2θ ≈ 41° also ap-
oxide types. Such difference is indeed a clear indication that the ox- pears in the diffraction patterns of the coatings. Its precise identifica-
ides were developed by two distinct and different processes (in-flight tion is hindered by its very low intensity; however, it is hypothesised
oxidation and post-deposition oxidation, respectively). In both cases, that it might belong to W (although its position does not match exact-
Al is oxidised preferentially; indeed, according to Ellingham's dia- ly with that reported in the JCPDF 4-806 file) or to (W,Cr)2C (accord-
gram, Al has greater high-temperature affinity for oxygen than Cr, ing to [32]). The formation of W would further corroborate to the
W and Fe. During in-flight oxidation, however, Al is almost the only occurrence of some decarburisation of WC during spraying, whereas,
oxidised element, whereas, during post-deposition oxidation, the if (W,Cr)2C were developed, this might suggest some interaction of
other metallic elements are also affected. dissolved WC with the metal matrix.
Presumably, WC particles were also affected during spraying: they Presumably, the fraction of W and C which remained dissolved within
experienced some decarburisation and dissolution in the molten ma- the matrix as a consequence of carbide dissolution (without entering the
trix. Microstructural evidence of these phenomena includes the mixed carbide phase (W,Cr)2C which might have appeared in the coat-
rounded edges of the carbide particles in the coatings (compare ings) further promoted its amorphisation during impact quenching (as
Fig. 5A, B to the original morphology of WC in the feedstock powder, discussed above), by opposing further hindrance to crystallisation [29].
Fig. 1C), the disappearance of some of the original sub-micrometric The two key microstructural features mentioned above, namely
WC particles (again, Fig. 5A, B can be compared to Fig. 1C), and the decarburisation and oxidation, were quantified by peak fitting on
formation of a few bright, W-rich matrix areas with no recognisable XRD patterns and by image analysis on SEM micrographs, respectively.
4084 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

Fig. 4. Cross-sectional BSE-SEM micrographs (1000 × magnification) of the HVOF-sprayed coatings deposited onto C40 steel plates: WC–FeCrAl #1 (A), #2 (B), #3 (C), #4 (D), #5
(E) and WC–CoCr reference (F).

Specifically, decarburisation was quantified through the index of coatings, however, can be considered meaningful, particularly because
carbide retention I, defined in accordance with [33] (Eq. (2)): the intensity of the peak at 2θ ≈ 41° varies between samples in the
same way as that of the W2C peaks, i.e. the coatings with lower amount
IWC of W2C also contain less W or less (W,Cr)2C.
I¼ : ð2Þ
I WC þ I W2C þ IW The results (Table 3) suggest that the degree of decarburisation is
primarily controlled by the oxygen/fuel ratio (λ, Table 2) during
Where: spraying; accordingly, highest decarburisation (lowest I value, sam-
ples #1 and #4) is associated to the lowest values of λ, whereas the
IWC integral intensity of the main WC diffraction peak at less decarburised samples (highest I) are those sprayed with highest
2θ = 35.6°; λ values (samples #2 and #3).
IW2C integral intensity of the main W2C diffraction peak at Quantification of oxide contents by image analysis (Table 3)
2θ = 39.8°; shows that, first of all, the WC–FeCrAl coatings are definitely more
IW integral intensity of the main W diffraction peak at oxidised than the WC–CoCr reference, consistent with the EDX re-
2θ = 40.3°. sults. This can also be qualitatively seen by comparing Fig. 4A–E to F
and Fig. 5A–B to C. Specifically, since the WC–FeCrAl #1 coating and
Integral intensities were obtained by fitting the diffraction peaks the WC–CoCr reference were deposited under identical process pa-
shown in Fig. 2 using pseudo-Voigt peak functions. Since the peak rameters using feedstock powders with similar particle size distribu-
at 2θ ≈ 41° cannot be undoubtedly ascribed to W, it was not employed tions (Table 1), the different oxide contents must have been caused
for the computation of I. This might cause I to be somewhat overesti- by higher oxidation tendency of the FeCrAl matrix compared to the
mated, since the amount of W possibly lost in the formation of W or CoCr one. This is probably due to the large reactivity between Al
of (W,Cr)2C is neglected. The ranking between the various WC–FeCrAl and oxygen at high temperature, as discussed previously.
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4085

on the key role of the oxygen-fuel ratio to determine properties such as


oxide content in HVOF-sprayed coatings were drawn in [34].

3.2. Microindentation testing

Indentation testing (Fig. 8) shows that the hardness of the WC–


FeCrAl coatings (roughly comprised between 1000 HV3N and 1100
HV3N) and their modulus (around 200 GPa) is lower than those of
the WC–CoCr reference (approximately 1200 HV3N and 350 GPa, re-
spectively). However, the mechanical properties of the WC–FeCrAl
coatings can be modified, to some extent, by the selection of process
parameters, just as it happened for the carbide retention and oxide
content discussed in Section 3.1. Specifically, Fig. 8 suggests that
hardness and modulus of the WC–FeCrAl coatings follow roughly in-
verse trends, i.e. samples exhibiting the highest hardness (samples
#4 and #5) possess lower modulus and vice versa.
Specifically, the differences between the modulus values of the
various WC–FeCrAl coatings are significantly large, compared to the
associated standard deviations (plotted in Fig. 8 as error bars), where-
as the differences between hardness values are quite small. It should
preliminarily be remarked that the standard deviations of the hard-
ness measurements, of about 70–90 Vickers units in all cases, are def-
initely consistent with the typical standard deviations associated to
the Vickers hardness values of HVOF-sprayed cermet coatings, as
seen e.g. in [35,36]. Due to the presence of scattered defects and inho-
mogeneities (such as oxide inclusions, weak interlamellar bound-
aries, etc.), indeed, hardness values measured on thermally sprayed
coatings usually exhibit significant scatter [36]. The actual statistical

Fig. 5. Cross-sectional BSE-SEM micrographs (details) of the WC–FeCrAl coatings #1


(A) and #5 (B) and of the WC–CoCr reference (C). In panel A: 1 = splat-boundary
oxide stringer; 2 = spherical intralamellar oxide inclusion; 3 = decarburised area
with no recognizable WC particles.

Another important feature highlighted by image analysis is the


possibility to vary, within certain limits, the oxide content of WC–
FeCrAl coatings by adjusting the process parameters (Table 3). The
lowest volume fraction of oxide inclusions is exhibited by samples
#4 and #5, which are also those with the lower oxygen content, as
determined by EDX micro-analysis (although such quantitative as-
sessment has important limitations, as mentioned previously). These
coatings were deposited using oxygen/fuel ratios closer to stoichiome-
try (Table 2), which implies less residual O2 in the flame to promote
in-flight and post-impact oxidation. Parameter set #1 did not produce
equally low oxidation in spite of its similarly low oxygen/fuel ratio be-
cause, since the flow rates of oxygen and kerosene are the highest, the
thermal power in the HVOF gas jet and the resulting heat flux into the
substrate are maximised: this presumably promoted post-deposition
oxidation at lamellar boundaries, according to the mechanisms outlined
previously.
Fig. 6. (A) EDX spectra of the metal matrix between WC particles and of the decar-
Fig. 7A effectively summarises the fact that the oxygen/fuel ratio (λ) burised area marked by label “3” in Fig. 5A, and (B) EDX spectra of oxides (splat-
is the primary, though not the only, factor controlling microstructural boundary oxide marked by label “1” and intralamellar oxide marked by label “2” in
features such as decarburisation and oxide content. Similar conclusions Fig. 5A) compared to the overall spectrum of the complete coating.
4086 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

Table 4
Two-sample Student's t-test on hardness values.

The null hypothesis corresponds to two populations having statistically identical aver-
age values.

Sample pair t Value Degrees of freedom Prob. > |t|

#1 vs. #2 − 0.41077 33 0.68389


#1 vs. #3 − 0.78002 31 0.44129
#1 vs. #4 − 4.64379 28 0.00007
#1 vs. #5 − 2.74808 30 0.01005
#1 vs. WC–CoCr − 8.79491 29 b 0.00001
#2 vs. #3 − 0.11684 36 0.90764
#2 vs. #4 − 2.63533 33 0.01271
#2 vs. #5 − 1.57368 35 0.12456
#2 vs. WC–CoCr − 5.42737 34 b 0.00001
#3 vs. #4 − 3.63208 31 0.00100
#3 vs. #5 − 1.97875 33 0.05624
#3 vs. WC–CoCr − 7.57162 32 b 0.00001
#4 vs. #5 1.43618 30 0.1613
#4 vs. WC–CoCr − 3.82645 29 0.00006
#5 vs. WC–CoCr − 5.05585 31 0.00002

tested populations of hardness values have identical average) is veri-


fied. This corroborates to the previous assumption that samples #4
and #5 are significantly harder than the other WC–FeCrAl samples.
Obviously, the probability of the “null hypothesis” also becomes neg-
ligibly small when comparing the hardness of all WC–FeCrAl coatings
to that of the WC–CoCr reference.
The behaviour of hardness and modulus data of WC–FeCrAl layers
is better clarified by plotting the results of micro-indentation tests
against the oxygen/fuel ratio λ (Fig. 7B). Although the statistical sig-
nificance of the differences between the average hardness values is
not guaranteed in all cases (as seen in Table 4), the graph shows
that hardness and modulus (just like the oxide content and carbide
retention index discussed in Section 3.1) primarily depend on λ:
they exhibit opposite linear dependence from λ, namely hardness de-
Fig. 7. (A) Microstructural features (carbide retention, oxide inclusions) and (B) micro-
creases while elastic modulus increases with increasing λ. When λ is
mechanical properties (micro-hardness, elastic modulus and H/E ratio assessed by
depth-sensing micro-indentation) as a function of the normalised oxygen-fuel ratio λ.
higher, indeed, coatings are generally less decarburised but more oxi-
dised. During the loading stage of depth-sensing indentation experi-
ments, oxides, especially those located at splat boundaries, are
significance of the differences between average hardness values was fractured (cracks propagating along splat boundaries, following
checked by applying the two-sample Student's t-test to all pairs of oxide stringers, are clearly seen in optical micrographs of Berkovich
hardness measurements (Table 4). The two-sample tests comparing micro-indentations, Fig. 9A), allowing inelastic interlamellar sliding.
the hardness values of coatings #1, #2 and #3 to those of coatings This results in larger indenter penetration and, consequently, lower
#4 and #5 return low probabilities (almost always around or below hardness. On the other hand, during the unloading stage, when the
0.05) that the “null hypothesis” (i.e. the hypothesis that the two elastic portion of deformation is recovered, the extremely large stiff-
ness of WC (modulus values > 500 GPa are usually found in the liter-
ature [37,38]) increases the modulus of the whole coating. Conversely,
coatings sprayed with low λ values are generally less oxidised, which
means better interlamellar cohesion and less irreversible inelastic de-
formation, but they are more decarburised, which means that the elastic
stiffness is decreased. The reason why the differences between hardness
values are less marked than those between modulus values (compared
to the respective standard deviations) may lie in the fact that scattered
defects (such as oxide inclusions at lamellar boundaries) affect hard-
ness more than modulus. Depending on the presence or absence of
such defects in the indented area, inelastic deformation during indenter
penetration can occur to significantly different extents, causing larger
scatter in hardness values, whereas the elastic recovery during unload-
ing is comparatively less affected, resulting in lower data scatter.
As a result of the trends outlined above, the H/E ratio between
hardness (converted in GPa) and elastic modulus also decreases line-
arly with λ (Fig. 7B). The parameter H/E indicates the ability of the
material to accommodate strains within elastic deformation regime,
without reaching the yield or failure limit [39,40]; it has important
Fig. 8. Vickers microhardness and elastic modulus values assessed by depth-sensing implications for the mechanical and tribological behaviours of coat-
micro-indentation on all HVOF-sprayed coatings deposited onto carbon steel substrate. ings and will be referred to in the forthcoming Sections 3.3 and 3.4
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4087

Fig. 10. Plots of εφψ as a function of sin2ψ acquired on WC–FeCrAl samples #1 and #5 at
φ = 0°, with linear fits to the experimental data points.

propagating along the weakest and most oxidised splat boundaries


(Fig. 9C).

3.3. Residual stress

The plane stress assumption underlying the use of Eq. (1) for X-
ray residual stress analysis is confirmed by the remarkably good line-
arity of the εφψ vs. sin 2ψ plots (two of which are shown in Fig. 10).
This indicates that no sub-surface shear stress gradients exist. As
such gradients are typically induced by surface machining (e.g. grind-
ing) [42], it may be assumed that the grinding and polishing opera-
tions performed on the present samples (as specified in Section 2.2)
did not alter significantly the stress state of the material. Manual
grinding with fine diamond papers and final polishing with diamond
slurries should indeed ensure that no significant near-surface defor-
mation is induced in the material. Accordingly, surface polishing of
thermally-sprayed coatings was also performed in [43,44] before X-
ray residual stress measurement. Such polishing operation is impor-
tant because the surface roughness of the as-deposited surface,
which is comparable to the penetration depth of X-rays in the mate-
rial, would significantly affect and alter the stress measurement [43].
All of the coatings (Table 5) exhibit compressive near-surface re-
sidual stresses. The values measured along the three φ directions do
not differ much from one another, which means that the near-
surface stress state is substantially equi-biaxial. The values are quite
consistent with those reported for other cermet coatings deposited
by liquid-fuelled HVOF torches [9,45–48]. These torches indeed con-
Fig. 9. Optical micrographs of Berkovich micro-indentations on the polished cross-
sections of WC–FeCrAl #3 (A) and WC–CoCr (B) coatings and of a high-load (10 N)
fer large kinetic energy to sprayed particles: compressive stresses
Vickers micro-indentation on the polished cross-section of sample WC–FeCrAl #3 are therefore caused by peening of previously deposited layers by
(C). The arrows indicate recognizable microcracks. new particles impinging at high velocity [9,49,50]. The fact that
deposition-related phenomena (i.e. peening effects), not the thermal
expansion mismatch between coating and substrate at the end of the
to assist in the interpretation of tribological behaviour and impact
behaviour.
Table 5
The WC–CoCr reference seems comparatively tougher than the Residual stresses in HVOF-sprayed cermet coatings. The elastic modulus values
WC–FeCrAl coatings, since Berkovich micro-indentations performed employed for the computation according to Eq. (1) are also shown in the table.
on the former are definitely less microcracked (compare Fig. 9B to
Coating/substrate E (GPa) σφ = 0° (MPa) σφ = 45° (MPa) σφ = 90° (MPa)
A). Quantification of the indentation fracture toughness of WC–
#1/C40 194 ± 22 − 260 ± 42 − 287 ± 44 − 288 ± 42
FeCrAl coatings, however, is unfeasible: attempts to produce high-
#2/C40 226 ± 16 − 219 ± 29 − 242 ± 24 − 273 ± 28
load, microcracked Vickers indentations usable for this purpose [41] #3/C40 260 ± 15 − 165 ± 22 − 160 ± 16 − 148 ± 20
were frustrated by the irregular cracking behaviour of the material. #4/C40 176 ± 11 − 241 ± 38 − 255 ± 36 − 305 ± 29
Instead of developing radial cracks propagating from the indentation #5/C40 203 ± 18 − 331 ± 43 − 360 ± 45 − 406 ± 48
corners, which are required for the assessment of indentation fracture #1/AA6082 194 ± 22 − 270 ± 47 − 291 ± 41 − 287 ± 43
WC–CoCr/C40 352 ± 23 − 415 ± 66 − 420 ± 49 − 453 ± 58
toughness, indeed, the WC–FeCrAl coatings exhibit multiple cracks
4088 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

deposition process, are responsible for the compressive stress state in moments and to quenching should be minimum. This sample, by con-
WC–FeCrAl coatings is confirmed by the substantial identity of the re- trast, is the one exhibiting the lowest compressive residual stress
sidual stress states measured on coating #1 deposited onto two sub- (Table 5).
strates with largely different thermal expansion coefficients (carbon This confirms that the thermal expansion mismatch between coat-
steel and aluminium alloy: Table 5). As mentioned in Section 2.1, in- ing and substrate does not have a major influence on the residual
deed, the deposition process was periodically interrupted in order to stress state of the present coatings, as observed previously, and that
avoid system overheating, which keeps stresses due to thermal ex- peening effects are largely preponderant over quenching stress
pansion mismatch down to a minimum. contributions.
It is important to remark that coating material subjected to peen-
ing was exposed onto the sample surface after the outermost layer
was removed by grinding and polishing operations: the very first sur- 3.3. Dry sliding wear behaviour
face layer, indeed, often has slightly tensile residual stress because of
the absence of peening, as shown e.g. in [45,48–50]. Specifically, The sliding wear rates measured on the various WC–FeCrAl coat-
micrometre measurements indicated that approximately 60 μm of ings after ball-on-disc tests (Table 6), roughly comprised between
material were removed from the coating surface after grinding 5 ∗ 10 − 8 mm 3/(Nm) and 1 ∗ 10 − 7 mm 3/(Nm), exhibit limited, yet
+ polishing. Unlike the thickness values assessed by image analysis significant differences. The best wear resistance is shown by
(listed in Table 3), which are referred to the mean surface line, micro- samples #4 and (particularly) #5: based on Sections 3.1–3.3, these
metre measurements are referred to the highest surface peaks. Since coatings possess lower oxide content, higher micro-hardness and
all of the coatings possess Rmax ≈ 45–50 μm (assessed through optical higher (H/E) ratio than the other WC–FeCrAl samples; sample #5
confocal profilometry), grinding and polishing took the original sur- also exhibits the largest compressive residual stress. All of these fac-
face roughness away entirely and further removed approximately tors probably concur to promote highest wear resistance. According-
10–15 μm of material from the coating surface, which roughly corre- ly, sample #3, which features low hardness, the lowest (H/E) ratio
sponds to the layer deposited during one torch cycle. The overall (Fig. 7B), high oxide content and the lowest compressive stress, also
thickness reduction of each coating was therefore of ≈35 μm. has the lowest wear resistance. This confirms that sliding wear resis-
Measurements performed under these experimental conditions tance depends on a combination of these factors. The substrate mate-
are certainly more significant than stress assessments on as- rial, on the other hand, has little or no influence on the sliding wear
deposited surfaces (where no peening has occurred), not only be- behaviour, as the wear rates of coating #1 deposited onto steel and al-
cause surface roughness may alter the stress measurement, as men- uminium substrates are very similar (Table 6).
tioned previously, but also because, in many practical applications In order to provide an explanation for the dependence of sliding
(especially those involving sliding wear contacts), HVOF-sprayed cer- wear on factors such as oxidation, hardness, H/E ratio and residual
met coatings are ground and polished in order to obtain smooth sur- stress, the wear mechanisms were investigated. All of the WC–FeCrAl
face finish. coatings experience identical wear mechanisms (Fig. 11). On the one
More specifically, sample #5 exhibits the largest compressive hand, single carbide particles are pulled out of the coating surface
stress (which is also qualitatively seen through the large slope of its (Fig. 11B, label 1), since abrasion of the surrounding metal matrix
εφψ vs. sin 2ψ graphs, Fig. 10): this is probably due to the fact that leaves these particles progressively unsupported (as seen e.g. in
the intensity of peening is maximised by the combined effects of rath- Fig. 11C). On the other hand, the formation and the propagation of
er large particle impact velocity (quite large oxygen and kerosene brittle cracks on the coating surface (as seen in Fig. 11D) lead to the
flow rates were employed, Table 2, which accelerates the particles removal of near-surface portions of material (Fig. 11B, label 2). Both
at large speeds [50,51]) and low oxide content (Section 3.1 and carbide pull-out and brittle cracking are well known sliding wear pro-
Table 3). The latter makes the deposited material more prone to be cesses for thermally-sprayed cermets (see [54–56] and [56–58], re-
plastically deformed and, therefore, peened by new impinging particles. spectively), but the latter is probably the one responsible for most
Differently from the properties examined in Sections 3.1 and 3.2, of the recorded wear loss. Brittle cracking is likely caused by the com-
therefore, residual stresses do not depend only (or primarily) on λ, bination of hertzian surface stresses (under ball-on-plane configura-
but on the combined effect of λ and of the total gas flow rate; the lat- tion, tensile radial stresses are produced around the edge of the
ter, indeed, has huge influence on in-flight particle velocity [50,51]. contact area) with frictional stresses: this combination indeed pro-
Accordingly, sample #4, sprayed not only with low λ value but also duces a maximum in the tensile normal stress at the trailing edge of
with low oxygen flow rate, exhibits lower compressive stress than the contact area [59], and the cyclic application of this stress at
sample #5. On the other hand, when oxidation becomes excessively every disc revolution, causing fatigue loading, increases the severity
large because of high λ values, the sensitivity of the material to peen- of the contact condition [60].
ing is reduced; sample #3 indeed exhibits the lowest compressive re- Large compressive residual stresses can clearly hinder the forma-
sidual stress of all WC–FeCrAl coatings. tion and propagation of cracks by reducing the magnitude of the
The thickness of the as-deposited coatings seems to have no obvi- overall surface tensile stresses. Moreover, when the H/E ratio is
ous relation with the residual stress state. In general, the thickness of
a thermally-sprayed coating can affect the residual stress state in two
ways. On the one hand, as a coating grows thicker, the bending mo- Table 6
Results of ball-on-disc dry sliding wear tests (sample wear rate), dry sand-steel wheel
ment, which arises during the final cooling because of the thermal ex- abrasion test (sample wear rate) and cyclic ball drop test (impact crater volume).
pansion mismatch with the substrate, becomes increasingly large.
This bending moment causes a stress gradient across the coating, giv- Sample Dry sliding Impact volume Abrasive
wear rate [mm3] wear rate
ing a tensile stress contribution near the top surface [52]. On the other −8
[∗10 3
mm /(Nm)] [∗10− 2 mm3/(Nm)]
hand, higher coating thickness implies (for a fixed number of torch
#1 C40 substrate 10.57 ± 0.97 0.244 1.53 ± 0.07
scans) larger thickness deposited per torch pass: this is usually be-
AA6082 substrate 11.84 ± 2.83 0.849
lieved to increase the tensile quenching stress contribution [53]. #2 7.12 ± 0.49 0.249 1.46 ± 0.21
Both of these factors, however, seem not to have significant effects #3 12.47 ± 2.43 0.203 1.57 ± 0.03
in the present case. For instance, based on the above considerations, #4 6.19 ± 1.75 0.175 1.63 ± 0.10
the thinnest coating (sample #3) should exhibit large compressive #5 4.94 ± 1.29 0.174 1.55 ± 0.06
WC–CoCr 3.94 ± 1.42 0.192 1.20 ± 0.08
residual stress, since tensile stress contributions due to bending
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4089

Fig. 11. BSE-SEM micrographs of the wear scar produced after ball-on-disc dry sliding wear testing on the WC–FeCrAl #3 coating. A: overview; B: higher magnification (label
1 = single carbide pull-out; label 2 = near-surface delamination of material); C: detail of matrix abrasion between WC particles; D: detail of microcracked area (the arrows indicate
a microcrack); and E: detail of a highly oxidised area with numerous microcracks (marked by arrows).

higher, the coating can accommodate larger near-surface strain be- contrast, “wavy” regions, clearly resulting from near-surface plastic
fore the elastic limit is reached [39,40]: since WC–FeCrAl coatings deformation, appear (Fig. 12B). The orientation of these features
are quite brittle (as shown by indentation testing in Section 3.2), (forming an angle with the plane of the original surface) indicates
cracking, rather than plastic deformation, occurs when the elastic that deformation is induced by shear stress acting on the coating sur-
limit is reached, so that higher H/E ratio eventually means lower ten- face. It is inferred that shear stresses caused by the tangential friction
dency to near-surface brittle fracture. Finally, brittle cracking is seen force [59] deform the ductile CoCr alloy matrix, which drags the fine
to be more frequent in areas where oxidation of the material is WC particles, giving rise to the observed undulations.
more pronounced (Fig. 11E: in the area shown in this micrograph, An important difference therefore emerges between the tribologi-
which exhibits many microcracks highlighted by arrows, a large num- cal behaviour of the WC–FeCrAl layers and that of the WC–CoCr refer-
ber of dark oxide inclusions can be seen), so that coatings with low ence. The WC–FeCrAl coatings fail like brittle materials, with surface
oxidation are less prone to cracking. cracks induced by the maximum tensile stress, as seen in Fig. 11B, D,
As a result, the sliding wear resistance of sample #5, which ex- and E, whereas the WC–CoCr coating follows the failure criterion for
hibits the most favourable combination between microstructural fea- ductile materials, with surface damage caused by shear stresses (the
tures, micromechanical properties and compressive residual stresses, well known Tresca failure criterion for ductile materials).
approaches that of the WC–CoCr reference (Table 6), although the This difference is explained by considering that, on the one hand,
wear mechanisms of the latter are somewhat different (Fig. 12). Ma- the value of the H/E ratio is not larger in the WC–CoCr coating (H/
trix abrasion (Fig. 12C) with single carbide pull-out occur on the WC– E = 3.8 ∗ 10 − 2, based on the data in Fig. 8) than in the WC–FeCrAl
CoCr layer as well; however, on a larger scale, the WC–CoCr coating layers (H/E comprised between 4.2 ∗ 10 − 2 and 6.9 ∗ 10 − 2, as shown
does not undergo cracking and near-surface removal of material. By in Fig. 7B). The ability of the WC–CoCr coating to accommodate
4090 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

3.4. Dry particle abrasion resistance

The wear rates produced in this test are much larger than those
found in sliding wear tests (Table 6), which indicates greater severity
of the particle abrasion tests over the sliding wear test.
In this case, the abrasive wear rates experienced by the various
WC–FeCrAl coatings do not differ significantly within experimental

Fig. 12. BSE-SEM micrographs of the wear scar produced after ball-on-disc dry sliding
wear testing on the WC–CoCr coating. A: Overview; B: higher magnification of an area
with numerous “wavy” features; and C: detail of matrix abrasion between WC particles.

surface strains without reaching the elastic limit is, therefore, not su-
perior to that of the WC–FeCrAl coatings, but, on the other hand,
when the elastic limit is reached, WC–FeCrAl layers crack almost im-
mediately, as discussed previously, whereas the WC–CoCr coating can
deform plastically without cracking, on account of its lower oxide
content and better toughness (qualitatively discussed in
Section 3.2). Its highly compressive residual stress state also helps
in mitigating the maximum surface tensile stress.
Though it is not central to the purpose of the present paper, it is
interesting to compare the present observations on the sliding wear
behaviour of the WC–CoCr coatings with previous studies by the
same authors on similar coatings. The near-surface plastic deform-
ability during dry sliding contacts indeed turns out to be a general
characteristic of WC–CoCr coatings deposited by liquid-fuelled HVOF
spraying. Identical “wavy” features related to surface plastic shearing,
without brittle cracking, were accordingly reported on coatings depos-
ited with GTV-K2 and Thermico-CJS torches [61,62]. This might be relat- Fig. 13. Secondary electron (SE)-SEM micrographs of the wear scar produced by dry
particle abrasion testing on the WC–FeCrAl #4 coating. A: Overview (label 1 = abrasive
ed to common features of coatings deposited by this kind of torches,
groove; label 2 = indentation; label 3 = surface crack); B: higher magnification view
including low oxide content, low decarburisation and compressive re- (label 1 = abrasive groove; label 3 = delaminated area); and C: detail of the area cir-
sidual stress state [9,10]. cled in panel B (arrows: chipping of the material along the side of the abrasive groove).
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4091

error (Table 6). All of the WC–FeCrAl samples experience a combina- shows a detail of the circled area of the groove labelled “1” in
tion of abrasive grooving (ploughing/cutting: label 1 in Fig. 13A, B), Fig. 13B. As deformation accumulates because of repeated abrasion
particle indentation (Fig. 13A, label 2) and brittle cracking. Some and indentation events, cracks grow on the surface and immediately
cracks are clearly seen in the overview micrograph of Fig. 13A (label below it, until an entire portion of material is delaminated. Accord-
3): they probably originate from the accumulation of elastoplastic de- ingly, a region where delamination of near-surface material occurred
formation in the material subjected to repeated abrasive grooving as a consequence of surface and sub-surface crack propagation is
and particle indentation. Cracking and chipping of the lateral ridges recognisable in Fig. 13B (label “3”). The morphology seen in Fig. 13B
of an abrasive groove are indeed seen in Fig. 13C (arrows), which is indeed remarkably similar to that produced by model experiments
on brittle coatings subjected to multi-cycle scratch tests (as shown
e.g. in [63,64]).
The WC–CoCr coating, by contrast, possesses somewhat better
abrasive wear resistance (Table 6). This sample indeed experiences
no obvious microcracking and no brittle removal of material
(Fig. 14A–C). The dominant wear mechanisms are abrasive grooving
(Fig. 14A, B, label “1”) and indentation of loose particles (Fig. 14A, B,
label “2”). Specifically, particle indentation seems much more fre-
quent, which is probably a consequence of the higher toughness
and ductility of this material, allowing penetration of the indenter
with limited or no cracking, in accordance with the results of indenta-
tion testing discussed in Section 3.2.
It would therefore seem that the resistance to cracking plays a more
important role in determining particle abrasion resistance than it does
in dry sliding wear conditions, where other factors (e.g. the H/E ratio,
the residual stress) are equally influential (as discussed in Section 3.3).
Toughness has accordingly been identified in the literature as an impor-
tant parameter in controlling particle abrasion resistance of brittle mate-
rials [65], including cermet coatings [56,66]. The lower oxide content,
better ductility and better resistance to interlamellar cracking of the
WC–CoCr sample are therefore decisive in determining its better abrasive
wear performance, whereas the slight differences existing between the
properties of the WC–FeCrAl samples are presumably too small to be ap-
preciated by this testing technique, which is probably not sensitive
enough.

3.5. Cyclic impact behaviour

The impact volumes of the various samples deposited onto the


steel substrate (Table 6) do not differ much from one another. The
largest difference occurs between the impact volumes of coating #1
deposited onto steel and onto aluminium: under the present impact
test conditions, deformation occurs in the coating as well as in the sub-
strate, so that the overall system deformation depends much on the
substrate properties. This means that all of the coatings deposited
onto the same substrate material are subjected to similar deformations.
Quite remarkably, the WC–FeCrAl coatings and the WC–CoCr refer-
ence exhibit very similar cracking behaviour, dominated by the forma-
tion of circumferential ring cracks along the periphery of the contact
region (Fig. 15). These ring cracks propagate from the coating surface
and extend across most of the thickness (Figs. 16A and 17A), although
they do not reach the substrate interface. No cracks appear inside the
coatings in the middle of the impact region (Figs. 16C and 17C), and no
delamination from the substrate is visible (Figs. 16A, C and 17A, C).
This cracking behaviour can be interpreted with reference to the spheri-
cal indentation models given by Chai and Lawn for hard coatings depos-
ited onto ductile substrates [67]. According to Chai and Lawn, through-
thickness cracking in the coating begins in the middle of the contact
area (radial cracks) when the radius of the contact region (a) is compa-
rable to the coating thickness (d), whereas it occurs preferentially
along the periphery of the contact area (circumferential ring cracks)
when d/a >0.1. In the present case, from optical profilometry measure-
ments on the impact mark, a ≈ 1.7 mm→ d/a ≈ 0.1, i.e. the present test
conditions are roughly at the boundary between indentation on inter-
mediate thickness coatings and indentation on thin coatings. Accord-
Fig. 14. Secondary electron (SE)-SEM micrographs of the wear scar produced by dry
particle abrasion testing on the WC–CoCr coating. A: overview (label 1 = abrasive
ingly, circumferential cracks occur preferentially, but they do not
groove; label 2 = indentation); B: higher magnification view (label 1 = abrasive extend across the entire coating thickness, as it would be expected in
groove; label 2 = indentation); and C: detail of abrasive grooves. the Chai and Lawn model for thin layers [67].
4092 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

The critical condition for the onset of circumferential ring cracking


given by Chai and Lawn for thin coatings is therefore applicable to the pre-
sent systems only with a certain degree of approximation; nonetheless, it
can provide very useful indications in order to interpret the results shown
in Figs. 15–17. According to this model [67], the critical load Pcrit for the
onset of circumferential ring cracking is (Eq. (3)):

ES 2
P crit ¼ A S a : ð3Þ
EC C

Fig. 15. Optical micrographs of the surfaces of the WC–FeCrAl #1 (A) and WC–CoCr (B)
coatings deposited onto steel substrate and of the WC–FeCrAl #1 coating deposited
onto AA6082 substrate (C), after cyclic ball drop testing. Fig. 16. Cross-sectional BSE-SEM micrographs of the WC–FeCrAl sample #1 deposited
onto C40 steel substrate, after cyclic ball drop test. A: Edge of the impact area, with
cross-sectional view of circumferential ring cracks; B: detail of a circumferential ring
crack; and C: middle of the impact area, with no visible cracks.
G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094 4093

This model is based on classical mechanic equations, not on fracture


mechanics; therefore it neglects the effects of defects, pores, inclusions,
etc. However, it is apparent from Eq. (3) that circumferential ring crack-
ing depends on the matching between the elastic modulus of the coat-
ing and that of the substrate, and on the SC/EC ratio of the coating. Since
the mechanical strength SC of a material is related to its hardness [68],
the SC/EC ratio depends on the H/E ratio of the coating.
Although the WC–FeCrAl coatings are more brittle and contain more
inclusions (e.g. oxide inclusions) than the WC–CoCr one, which may fa-
vour the onset of cracking, their lower elastic modulus (Fig. 8) is better
matched to that of the substrate materials (ES ≈ 210 GPa for steel,
ES ≈ 70 GPa for aluminium alloys [69]) and their H/E ratio is larger. Con-
sequently, the circumferential cracking behaviour of all cermet coatings
(all subjected to analogous deformations, as noted previously) is
similar.
It is also observed that the deformation of the coated systems dur-
ing ball drop induces normal stresses which are mainly parallel to the
coating / substrate interface [67], i.e. these stresses lay in the plane of
the lamellae and are directed parallel, not perpendicular, to the inter-
lamellar oxide stringers. Consequently, these interlamellar oxides are
not much affected by the deformation-induced stresses. Accordingly,
circumferential cracks have rather straight propagation paths across
the thickness of WC–FeCrAl coatings, without branching along inter-
lamellar oxides, as seen in Fig. 16B. Interestingly, higher degree of
crack branching occurs in the WC–CoCr coating (Fig. 17B). The cir-
cumferential cracks in this coating indeed develop numerous second-
ary branches which follow quite tortuous propagation paths: in low
magnification micrographs, this appears as an apparent increase in
porosity at the boundary of the impact region (Fig. 17A). This might
be ascribed to the greater stiffness of this coating, which tends to de-
velop larger stresses for a given imposed deformation.

4. Conclusions

This research focused on the characterisation of WC–FeCrAl cer-


met coatings deposited by liquid-fuelled HVOF-spraying onto steel
and aluminium alloy substrates. Their microstructural features, micro-
mechanical properties, sliding and abrasive wear behaviour and cyclic
impact resistance were studied as a function of the process parameters
and were compared to the characteristics of a reference WC–CoCr layer,
manufactured using the same technique.
The main conclusions of this study are as follows:

• Deposition mechanisms of WC–FeCrAl powder particles involve in-


flight melting and homogenisation of the metal matrix, in-flight and
post-deposition interaction with oxygen, as well as some decarbur-
isation and dissolution of WC. Decarburisation decreases and oxida-
tion increases as the oxygen/fuel ratio (λ) increases.
• Micro-hardness, elastic modulus and H/E ratio also vary with λ (low
λ = higher hardness, lower modulus, higher H/E ratio), because
these properties depend on oxidation and decarburisation.
Fig. 17. Cross-sectional BSE-SEM micrographs of the WC–CoCr coating deposited onto • Compared to the WC–CoCr reference, the WC–FeCrAl layers are
C40 steel substrate, after cyclic ball drop test. A: edge of the impact area, with cross- more oxidised; they possess lower hardness, lower modulus, but
sectional view of circumferential ring cracks; B: detail of branching in a circumferential higher H/E ratio. Higher tendency towards interlamellar cracking
ring crack; and C: middle of the impact area, with no visible cracks. is also qualitatively noted in WC–FeCrAl coatings.
• The WC–FeCrAl coatings possess compressive residual stresses,
Where: whose magnitude seems to depend both on λ and on the total gas
flow rate during spraying. WC–CoCr is also under compression.
ES, EC elastic modulus of the substrate and of the coating • Small but significant differences exist between the sliding wear
(respectively); rates of the various WC–FeCrAl coatings. Sliding wear resistance is
SC mechanical strength of the coating; enhanced by a combination of high compressive residual stresses,
A proportionality coefficient (in hertzian contacts, A = 2π / low oxide content, high H/E ratio. The sliding wear performance
(1 − 2υS) with υS = Poisson's ratio of the substrate, but of the best WC–FeCrAl coating is comparable to that of the WC–
this formulation is not applicable to the present case, CoCr reference. WC–FeCrAl layers are therefore a good alternative
where significant plastic deformation of the substrate to WC–CoCr when sliding wear is involved (e.g. protection of shafts,
occurs). bearings, seals, piston sleeves, etc.).
4094 G. Bolelli et al. / Surface & Coatings Technology 206 (2012) 4079–4094

• In dry particle abrasion, by contrast, no WC–FeCrAl coating approaches [22] A. Portinha, V. Teixeira, J. Carneiro, M.G. Beghi, C.E. Bottani, N. Franco, R. Vassen,
D. Stoever, A.D. Sequeira, Surf. Coat. Technol. 188–189 (2004) 120.
the performance of WC–CoCr, as their abrasive wear resistance is im- [23] J.F. Archard, J. Appl. Phys. 24 (8) (1953) 981.
paired by excessive oxidation-induced brittleness. WC–FeCrAl coatings [24] G. Bolelli, V. Cannillo, L. Lusvarghi, S. Riccò, Surf. Coat. Technol. 200 (9) (2006)
are less suitable for operation under particle abrasion conditions (e.g. 2995.
[25] G. Bolelli, V. Cannillo, L. Lusvarghi, T. Manfredini, Wear 261 (11–12) (2006) 1298.
mills, crushers, conveyors for particulate materials, etc.). [26] M. Barletta, G. Bolelli, B. Bonferroni, L. Lusvarghi, J. Therm. Spray Technol. 19
• The cyclic impact behaviour of the WC–FeCrAl coatings does not (1–2) (2010) 358.
differ from that of the WC–CoCr reference. All coatings exhibit cir- [27] L.-M. Berger, S. Saaro, T. Naumann, M. Wiener, V. Weihnacht, S. Thiele, J.
Suchánek, Surf. Coat. Technol. 202 (18) (2008) 4417.
cumferential ring cracking with no interface delamination. Higher [28] S. Thiele, K. Sempf, K. Jaenicke-Roessler, L.-M. Berger, J. Spatzier, J. Therm. Spray
H/E ratio and lower elastic modulus compensate the greater brittle- Technol. 20 (1–2) (2011) 358.
ness of WC–FeCrAl by reducing stress concentration in the coating [29] C. Verdon, A. Karimi, J.-L. Martin, Mater. Sci. Eng., A 246 (1–2) (1998) 11.
[30] G. Espie, A. Denoirjean, P. Fauchais, J.C. Labbe, J. Dubsky, O. Schneeweiss, K.
during impact events.
Volenik, Surf. Coat. Technol. 195 (1) (2005) 17.
[31] S. Deshpande, S. Sampath, H. Zhang, Surf. Coat. Technol. 200 (18–19) (2006)
Further developments of this research will include an assessment
5395.
of the corrosion resistance of WC–FeCrAl coatings deposited onto dif- [32] P. Stecher, F. Benesovsky, H. Nowotny, Planseeber. Pulvermet. 12 (1964) 89 (in
ferent substrate materials and on their ionic release in liquid media of German).
interest for possible applications to the food processing industry (as [33] C. Bartuli, T. Valente, F. Cipri, E. Bempord, M. Tului, J. Therm. Spray Technol. 14 (2)
(2005) 187.
mentioned in the Introduction). [34] J. Saaedi, T.W. Coyle, H. Arabi, S. Mirdamadi, J. Mostaghimi, J. Therm. Spray Tech-
nol. 19 (3) (2010) 521.
Acknowledgements [35] M. Factor, I. Roman, J. Therm. Spray Technol. 11 (4) (2002) 468.
[36] T. Valente, Surf. Coat. Technol. 90 (1997) 14.
[37] M. Lee, S. Gilmore, J. Mater. Sci. 17 (1982) 2657.
The authors are grateful to Mr. Ferri (Parma Spray Italia S.r.l.) for [38] H. Warlimont, in: W. Martienssen, H. Warlimont (Eds.), Springer Handbook of
his precious assistance during coating deposition. Many thanks to Condensed Matter and Materials Data, Springer, Berlin, Germany, 2005, p. 458.
[39] A. Leyland, A. Matthews, Wear 246 (2000) 1.
Dr. Ing. Benedetta Bonferroni for her help with the experimental ac- [40] G.W. Stachowiak, A.W. Batchelor, Engineering Tribology, Second Edition, Butter-
tivities and to Dr. Miriam Hanuskova (University of Modena and Reg- worth-Heinemann, Woburn, MA, USA, 2001, p. 464.
gio Emilia) for the particle size distribution measurement. [41] K. Niihara, R. Morena, D.P.H. Hasselman, in: R.C. Bradt (Ed.), Surface Flaws, Statis-
tics and Microcracking, Plenum Press, New York, 1983, p. 97.
The research was funded by Centro Interdipartimentale INTERMECH
[42] H. Dölle, J.B. Cohen, Metall. Trans. A 11 (1) (1980) 159.
MO.RE. located at the Faculty of Engineering “Enzo Ferrari”, University [43] V. Teixeira, M. Andritschky, W. Fischer, H.P. Buchkremer, D. Stöver, Surf. Coat.
of Modena and Reggio Emilia and by Regione Emilia Romagna, Italy. Technol. 120–121 (1999) 103.
[44] J. Stokes, L. Looney, J. Mater. Eng. Perform. 18 (1) (2009) 21.
[45] R.C. Souza, W.L. Pigatin, Corros. Rev. 21 (1) (2003) 75.
References [46] S. Houdková, M. Kašparová, F. Zahálka, J. Therm. Spray Technol. 19 (5) (2010)
893.
[1] B.S. Mann, V. Arya, P. Joshi, J. Mater. Eng. Perform. 14 (4) (2005) 487. [47] R. Ahmed, M. Hadfield, J. Therm. Spray Technol. 11 (3) (2002) 333.
[2] F. Rastegar, D.E. Richardson, Surf. Coat. Technol. 90 (1997) 156. [48] R. Gadow, A. Candel, M. Floristán, Surf. Coat. Technol. 205 (2010) 1074.
[3] B.S. Mann, V. Arya, A.K. Maiti, M.U.B. Rao, P. Joshi, Wear 260 (2006) 75. [49] P. Bansal, P.H. Shipway, S.B. Leen, Acta Mater. 55 (15) (2007) 5089.
[4] D.W. Wheeler, R.J.K. Wood, Wear 258 (2005) 526. [50] S. Kuroda, Y. Tashiro, H. Yumoto, S. Taira, H. Fukanuma, S. Tobe, J. Therm. Spray
[5] P. Chivavibul, M. Watanabe, S. Kuroda, M. Komatsu, Surf. Coat. Technol. 202 Technol. 10 (2) (2001) 367.
(2008) 5127. [51] J.A. Picas, M. Punset, M.T. Baile, E. Martín, A. Forn, Surf. Coat. Technol. 205 (2011)
[6] L. Moskowitz, K. Trelewicz, J. Therm. Spray Technol. 6 (3) (1997) 294. S364.
[7] S.J. Matthews, B.J. James, M.M. Hyland, Mater. Charact. 58 (2007) 59. [52] A.M. Kamara, K. Davey, J. Mech. Eng. Sci. 222 (2008) 2053.
[8] M. James, Sealing Technol. 2004 (10) (2004) 9. [53] J. Matejicek, S. Sampath, D. Gilmore, R. Neiser, Acta Mater. 51 (2003) 873.
[9] D.E. Crawmer, in: J.R. Davis (Ed.), Handbook of Thermal Spray Technology, ASM [54] J.M. Guilemany, S. Dosta, J.R. Miguel, Surf. Coat. Technol. 201 (2006) 1180.
International, Materials Park, OH, USA, 2004, p. 54. [55] Q. Yang, T. Senda, A. Ohmori, Wear 254 (2003) 23.
[10] H. Herman, S. Sampath, R. McCune, Mater. Res. Soc. Bull. 25 (7) (2000) 17. [56] Y. Qiao, T.E. Fischer, A. Dent, Surf. Coat. Technol. 172 (1) (2003) 24.
[11] H.C. Starck, AMPERIT thermal spray powder brochure, http://www.hcstarck.com/ [57] S. Usmani, S. Sampath, D.L. Houck, D. Lee, Tribol. Trans. 40 (3) (1997) 470.
hcs-admin/file/ae23e4b2316f58850131b30f364743f9.de.0/AMPERIT_Inlet. [58] V. Stoica, R. Ahmed, T. Itsukaichi, Surf. Coat. Technol. 199 (2005) 7.
pdf(accessed 07/01/2012). [59] G.M. Hamilton, Proc. Inst. Mech. Eng. C: Mech. Eng. Sci. 197 (1983) 53.
[12] “Commission Regulation (EC) No 790/2009: amending, for the purposes of its adap- [60] G.W. Stachowiak, A.W. Batchelor, Engineering Tribology, Second Edition, Butter-
tation to technical and scientific progress, Regulation (EC) No 1272/2008 of the Eu- worth-Heinemann, Woburn, MA, USA, 2001, p. 571.
ropean Parliament and of the Council on classification, labelling and packaging of [61] G. Bolelli, M. Bonetti, B. Bonferroni, L. Lusvarghi, L.-M. Berger, F.-L. Toma, R.
substances and mixtures”, Official Journal of the European Union 5.9.2009, L235/1. Putschmann, Proceedings of the 25th International Conference on Surface Modifica-
Available on-line at: http://eur-lex.europa.eu/LexUriServ/LexUriServ.do?uri=OJ:L: tion Technologies (STM-25), June 20–22 2011, (Trollhättan, Sweden). Available on-
2009:235:0001:0439:EN:PDF (last accessed 07-01-2012) line at:, http://www.ptw.hv.se/extra/measurepoint/?module_instance=3&name=
[13] S. Zimmermann, B. Gries, J. Fischer, in: E. Lugscheider (Ed.), Thermal Spray 2008: G.Bolelli&url=/dynamaster/file_archive/110812/8bd103a7bb2b893bc7388fd722
Crossing Borders: Proceedings of the International Thermal Spray Conference 5e70e7/Bolelli%5fWC%2d%28W%2cCr%292C%2dNi%5f21%2d06%2d2011%5f
2008, DVS Deutscher Verband für Schweißen, Düsseldorf, 2008. Session%204.pdf.
[14] U.S. Department of Health and Human Services, Cobalt–Tungsten Carbide: Powders [62] E. Bergonzini, G. Bolelli, B. Bonferroni, L. Lusvarghi, T. Varis, U. Kanerva, T.
and Hard Metals, 12th Report on Carcinogens, June 10 2011 Available on-line at:, Suhonen, J. Oksanen, O. Söderberg, S.-P. Hannula, Proceedings of the International
http://ntp.niehs.nih.gov/?objectid=03C9AF75-E1BF-FF40-DBA9EC0928DF8B15 Thermal Spray Conference 2011, DVS Deutscher Verband für Schweißen, Düssel-
(last accessed 07-01-2012). dorf, 2011, session: Wear Protection 2.
[15] A. Scrivani, N. Antolotti, S. Bertini, R. Groppetti, T. Gutema, A. Mangia, S. [63] Y. Xie, H.M. Hawthorne, Wear 233–235 (1999) 293.
Mucchino, G. Viola, in: C. Coddet (Ed.), Thermal Spray: Meeting the Challenges [64] H.M. Hawthorne, Y. Xie, Meccanica 36 (2001) 675.
of the 21st Century, ASM International, 1998, p. 1019. [65] K. Kato, K. Adachi, in: B. Bhushan (Ed.), Modern Tribology Handbook, Volume
[16] F.-W. Bach, T. Duda, M. Berthold, Schweiss. Schneid. 53 (1) (2001) 20 (in German). One, CRC Press, Boca Raton, FL, USA, 2001, p. 273.
[17] G. Bolelli, B. Bonferroni, J. Laurila, L. Lusvarghi, A. Milanti, K. Niemi, P. Vuoristo, [66] S.F. Wayne, S. Sampath, J. Therm. Spray Technol. 1 (4) (1992) 307.
Wear 276–277 (2012) 29. [67] H. Chai, B.R. Lawn, J. Mater. Res. 19 (6) (2004) 1752.
[18] S. Zimmermann, B. Gries, B. Brüning, Proceedings of the International Thermal [68] D. Tabor, The Hardness of Metals, Oxford University Press, Oxford, UK, 1951, p. 19.
Spray Conference 2011, DVS Deutscher Verband für Schweißen, Düsseldorf, [69] F. Goodwin, S. Guruswamy, K.U. Kainer, C. Kammer, W. Knabl, A. Koethe, G.
2011, session: Powders, Wires & Suspensions. Leichtfried, G. Schlamp, R. Stickler, H. Warlimont, in: W. Martienssen, H.
[19] W.C. Oliver, G.M. Pharr, J. Mater. Res. 7 (6) (1992) 1564. Warlimont (Eds.), Springer Handbook of Condensed Matter and Materials Data,
[20] V. Hauk, Structural and Residual Stress Analysis by Nondestructive Methods, Else- Springer, Berlin, Germany, 2005, p. 161.
vier, Amsterdam, The Netherlands, 1997, p. 17.
[21] J.K.N. Murthy, D.S. Rao, B. Venkataraman, Wear 249 (7) (2001) 592.

Das könnte Ihnen auch gefallen