Sunteți pe pagina 1din 18

Wear 261 (2006) 12981315

Wear behaviour of thermally sprayed ceramic oxide coatings


Giovanni Bolelli, Valeria Cannillo, Luca Lusvarghi , Tiziano Manfredini
Dipartimento di Ingegneria dei Materiali e dellAmbiente, Universit` di Modena e Reggio Emilia, Via Vignolese 905, 41100 Modena, MO, Italy a Received 6 July 2005; received in revised form 9 February 2006; accepted 10 March 2006 Available online 18 April 2006

Abstract The wear resistance of plasma-sprayed ceramic coatings (Al2 O3 , Al2 O3 13%TiO2 , Cr2 O3 ) has been investigated through pin-on-disk and dry sand-steel wheel tests, has been correlated to microstructural and micromechanical characteristics (microhardness, fracture toughness) and has been compared to well-known platings (such as Cr electroplating and electroless Ni) and HVOF-sprayed cermets (WC17%Co, WC10%Co4%Cr). Plasma-sprayed ceramics are hard but brittle: dry particles abrasion occurs through splats detachment. The toughest coating (Al2 O3 ) displays the highest wear resistance, which in fact overcomes HVOF-sprayed cermets and Cr electroplating, when a low number of wheel revolutions are considered. In pin-on-disk tests, no coating undergoes wear loss against the 100Cr6 ball, that possess lower hardness. Against the alumina ball, Al2 O3 and Al2 O3 TiO2 coatings show high wear rates and friction coefcients (due to chemical afnity), while Cr2 O3 possesses better wear resistance, lower friction coefcient and inicts less wear on the counterpart. Cr2 O3 wear scar consists in plastically deformed splats and debris forming a quite adherent protective tribolm. 2006 Elsevier B.V. All rights reserved.
Keywords: Sliding wear; Abrasive wear; Plasma spraying; Ceramic coatings; Friction coefcient

1. Introduction Thermal spraying is often considered as a potential alternative to traditional coating manufacturing techniques (such as hard chrome electroplating) for the production of wear-resistant coatings [14]. In fact, a large variety of hard materials (including ceramics and cermets) can be deposited on a cold (or moderately pre-heated) substrate [5,6], thus obtaining very hard coatings while preventing thermal alteration of the substrate itself (which invariably occurs in other hardfacing processes, such as in welding processes), which is a key requirement when design tolerances must be satised, thin-walled components are being considered, or heat-sensitive materials (like Al and Mg alloys) are being processed. Among the various thermal spraying techniques, plasmaspraying and HVOF-spraying are the most suitable for the production of high-quality wear resistant coatings and they are both technologically mature processes. HVOF was developed to overcome plasma-spraying limits: it is well known that HVOFspraying gives far superior results than plasma-spraying when

Corresponding author. Fax: +39 0592056243. E-mail address: lucalusv@unimore.it (L. Lusvarghi).

manufacturing cermet coatings, because of the much higher gas jet velocity and lower ame temperature resulting in coatings with extremely low porosity, low splats oxidation and low carbide decomposition and/or dissolution [7]; in fact, many experimental work concerning HVOF-sprayed cermet coatings tested under various conditions exist [813]. However, HVOF has its own limits: it produces high quality metallic alloy and cermet coatings, but these powders are difcult to prepare and very expensive. HVOF sprayed ceramic oxide coatings are still under investigation, but they are not widely used yet, because they must be tested and can be deposited just with few of the commercial torches [14,15]. Thus, plasma spraying is still the most widespread production technique for ceramic coatings, like Al2 O3 , Cr2 O3 . Such coatings are more porous and brittle than HVOF-sprayed cermets, because of intrinsic porosity of plasma-sprayed coatings due to lower particles in-ight velocity and quenching-induced microcracking in the ceramic splats [6,16]. Anyway, they possess very high hardness, due to their purely ceramic nature, they are almost insensitive to many corrosive environments and can stand high temperatures [17]. Thus, if the involved application does not require a liquid tight coating, but just wear resistance, plasma-sprayed ceramic oxide coatings can be a good solution. Besides, manufacturing of atmospheric plasma-sprayed ceramic coatings can be less

0043-1648/$ see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.wear.2006.03.023

G. Bolelli et al. / Wear 261 (2006) 12981315

1299

expensive than HVOF-sprayed cermets in many applications, since powder and processing costs are often lower [18,19], even though exceptions may exist and deposition efciency is also a concern. Therefore, many industrial processes make use of plasma-sprayed ceramic coatings, whose reproducibility is good once the optimal set of parameters has been found. For example, the food and medicine packaging industry does not only need wear resistance, but also the absence of heavy metals contamination: Al2 O3 and Al2 O3 TiO2 are often used for this reason in that eld. Therefore, plasma-sprayed hard ceramic coatings are still studied nowadays [2022]. Therefore, a thorough study of the wear resistance of thermally sprayed coatings must involve plasma-sprayed ceramics, which could represent an economical alternative to HVOFsprayed cermets in some industrial applications. Much research related to the basic wear mechanisms of plasma-sprayed oxides exists, since such coatings have been studied for a long time [2327]; however, there exist a few works comparing them to the characteristics of other thermally sprayed coatings as well as to other industrially widespread wear resistant coatings, such as hard chrome electroplating and nickel electroless plating [28,29]. Furthermore, to fully assess the industrial applicability of thermally sprayed coatings, and of plasma sprayed oxides in particular, wear maps should be experimentally obtained, as it is currently being done for massive sintered ceramics [3032]. Therefore, the aim of this study is both to provide an experimental assessment of the wear rates, wear mechanisms, friction coefcients of various plasma-sprayed ceramic coatings under different sliding wear conditions through pin-on-disk testing, as a rst step towards wear mapping of such materials, and to compare the performance of plasma-sprayed ceramic coatings, HVOF-sprayed cermets and metal platings under various wear conditions, in particular under dry sliding conditions and under dry particles abrasion conditions. 2. Materials and characterizations Three plasma-sprayed ceramic coatings, namely Al2 O3 (powder: Sulzer-Metco 105SFP, 31 + 3.5 m), Al2 O3 13%TiO2 (powder: Sulzer-Metco 130, 53 + 15 m) and Cr2 O3 (powder: Saint-Gobain #3033, 15 + 5 m), with a NiCoCrAlY (powder: Sulzer-Metco 461NS, 150 + 22 m) bond coat to improve adhesion, and two HVOF-sprayed cermet ones, namely WC17%Co (powder: Tafa 1343, 45 + 15 m, agglomerated and sintered) and WC10%Co4%Cr (powder: Sulzer Metco 5847, 53 + 11 m, agglomerated and
Table 1 Plasma spraying operating parameters Parameters Spray distance (mm) Cooling gas type and pressure Plasma gases ow rates Current (A) voltage (V) = power (kW) Feeding disk revolution speed (rpm) Carrier gas type and ow Al2 O3 105 Ar, 8 bar Ar: 50 slpm H2 : 15 slpm 580 68 = 39.44 12 Ar, 3.5 slpm

sintered), have been manufactured onto C40 steel plates (100 mm 100 mm 5 mm). The crystalline phases of all the three ceramic oxide powders have been studied with X-rays diffraction (XRD, Philips PW3710, Cu K radiation). The substrates were grit-blasted with 500 m alumina grits (SulzerMetco Metcolite-C) in a vacuum-operated blasting machine (Norblast) prior to coating deposition. Plasma-sprayed coatings were manufactured with a Sulzer-Metco F4-MB plasma torch, operated in Air Plasma Spraying (APS) mode in a C.A.P.S. plant (Centro Sviluppo Materiali S.p.A., Roma, Italy, co-shared with Universit` la Sapienza, Roma, Italy). The spraying parama eters for plasma-sprayed oxides are listed in Table 1, while spraying runs of cermet coatings have been described elsewhere [29]. Hard chrome electroplating and Ni14%P electroless plating on C40 steel have also been studied for reference. Both Cr electroplatings (on ground substrates) and NiP electroless platings (on micro-grit blasted substrates) were industrially manufactured by proprietary processes. The latter was supplied in the as-plated state, without thermal treatment: a heat treatment (5 C/min heating, 400 C treatment temperature for 1h, slow cooling inside the kiln) was performed on some of the samples (hereafter referred to as NiP tt). The treatment conditions were chosen following literature indications in order to achieve the highest possible hardness increase through precipitation of nanostructured NiP crystals [33]. The plasma sprayed coatings microstructure was determined by XRD and by scanning electronic microscopy (SEM, Philips XL30) on polished cross-sections (mounted in resin, ground with 400, 800, 1000, 2000 mesh SiC papers and polished with 3 and 0.5 m polycrystalline diamond paste). Image analysis was also performed on SEM images to determine coating porosity (UTHSCSA Image Tool v. 3.0 software). Roughness measurement was performed by optical prolometry (ConScan prolometer, CSM Instruments, Switzerland), determining the Ra and Rz parameters (UNI ISO 4287-1). A depth-sensing Vickers microindenter (CSM Instruments, Switzerland) was employed to measure Vickers microhardness, fracture toughness and elastic modulus on coatings polished cross-sections. Vickers microhardness was calculated on all coatings by measuring the indentations diagonals (1 N load, 15 s loading time) from SEM micrographs. Fracture toughness was determined on thermally sprayed coatings by high-load, cracked Vickers microindentations produced with 10 N load (5 N for Al2 O3 13%TiO2 due to its lower hardness and toughness), measuring the indentation diagonals and crack lengths from SEM micrographs and employing the

Al2 O3 13%TiO2 105 Ar, 8 bar Ar: 50 slpm H2 : 13 slpm 690 65 = 44.85 12 Ar, 3.5 slpm

Cr2 O3 105 Ar, 8 bar Ar: 45 slpm H2 : 15 slpm 650 67 = 43.55 15 Ar, 2.5 slpm

Bond coat 105 Ar, 8 bar Ar: 50 slpm H2 : 15 slpm 560 69 = 38.64 10 Ar, 2.7 slpm

1300 Table 2 pin on disk testing conditions Test 1 Sliding speed (m/s) Sliding distance (m) Normal load (N) Pin material E (GPa) 0.1 250 1 100Cr6 210 0.3 Test 2 0.1 250 1 Alumina 380 0.23

G. Bolelli et al. / Wear 261 (2006) 12981315

Test 3 0.1 250 5 100Cr6 210 0.3

Test 4 0.1 250 5 Alumina 380 0.23

Test 5 0.2 250 1 100Cr6 210 0.3

Test 6 0.2 250 1 Alumina 380 0.23

Test 7 0.2 250 5 100Cr6 210 0.3

Test 8 0.2 250 5 Alumina 380 0.23

EvansWilshaw formula [34]: KIC = 0.079(P/a3/2 ) log(4.5a/c) (2.1)

where KIC = MPa m0.5 , a the half diagonal of the indentation ( m), c the crack length ( m), and P is the load (mN). As reported in literature [35], the limitation for the use of this formula is that the ratio between the crack and the half diagonal length must be between 0.6 and 4.5. This condition has been veried in all the tests. Even if the formula has been developed for halfpenny-shaped cracks, it has been demonstrated that it is valid also for Palmqvist cracks [36]. The employment of this formula can already be found in literature for plasma spray ceramic coatings [37] and HVOF cermets [38]. Elastic modulus was determined on all coatings from the unloading part of instrumented indentation loading-unloading curves, by the OliverPharr formula [39]. The Poissons ratio (required for elastic modulus calculation) was assumed to be 0.23 for all ceramics and 0.3 for metals (Cr, NiP) and cermets. Following preliminary tests (which highlighted some effect of the indentation load on the elastic modulus value, especially for plasma-sprayed ceramics), a 5 N indentation load was chosen, with 5 N/min loading rate, 4 N/min unloading rate, 15 s loading time. A minimum of 20 indentations was performed for each hardness, toughness and elastic modulus measurement. Dry particles abrasion resistance was tested on all thermally sprayed coatings in their as-sprayed conditions and on as-plated hard chrome for reference (the measurement was not performed on Ni plating due to its low thickness causing the substrate to be quickly exposed) by a simple dry sand-steel wheel test, using
Table 3 Pin on disk testing plan (EHC = electrolytic hard chrome) Test 1 As-sprayed Al2 O3 As-sprayed Al2 O3 TiO2 As-sprayed Cr2 O3 As-sprayed WCCo As-sprayed WCCoCr As-plated EHC As-plated NiP As-plated NiP tt Polished Al2 O3 Polished Al2 O3 TiO2 Polished Cr2 O3 Polished WCCo Polished EHC Test 2 Test 3

a 200.1 mm diameter Fe360A steel wheel rotating at 75 rpm, FEPA 80 alumina grains (180 m average particle diameter) as abrasive medium with a 1 g/lap mass ux, and applying a 40.2 N normal load (Ceramic Instruments AP/87 abrasimeter). This test is a modied version of ASTM G65, with the main differences consisting in the use of steel wheel instead of a rubber wheel and of corundum in place of Ottawa sand. Ball-on-disk dry sliding tests were performed with a pin-ondisk tribometer (CSM Instruments, Switzerland) using 100Cr6 balls (manufacturers nominal hardness: 7 GPa; radius = 3mm) and sintered alumina balls (manufacturers nominal hardness: 19 GPa; radius = 3 mm) on (22 mm 22 mm 5 mm) coated plates obtained by cutting the (100 mm 100 mm 5 mm) plates. Plates are xed onto a rotating disk: the ball is xed into a steady ball holder pressed against the sample surface with a normal load (rotating unidirectional sliding). Eight different parameter sets have employed, varying the sliding speed, the normal load and the counterpart material. The parameters of the various tests (labelled tests 18) are listed in Table 2. For each test, the friction coefcient was measured on-line by the instrument; the wear rate of the ball was determined by measuring the worn cap diameter with an optical microscope and the wear rate of the sample was determined by measuring the area of the wear track cross-section by optical prolometry (ConScan prolometer, CSM Instruments, Switzerland: each area value is the average of four measurements) and calculating the resulting wear volume. All tests (from tests 1 to 8) were performed on plasma-sprayed ceramic coatings (Table 3): these data will provide important information on wear mechanisms and will serve as a basis for wear mapping of such coatings. A reduced set of

Test 4

Test 5

Test 6

Test 7

Test 8

G. Bolelli et al. / Wear 261 (2006) 12981315

1301

tests was performed on cermets and on other platings (namely test 3, 4, 7, 8; Table 3): they provide a good amount of data to evaluate the cermets sliding behaviour and are of use as reference values to compare the different performances. To assess the effect of surface nishing on the friction and wear behaviour, tests 7 and 8 have also been performed on polished plasmasprayed ceramics, HVOF-sprayed WC17%Co and electrolytic hard chrome (Table 3). These tribological samples were, in fact, laboratory polished using diamond abrasive papers with progressively smaller diamond particles, down to 2 m. 3. Results 3.1. Microstructure and micromechanical properties SEM micrographs in Fig. 1AF describe the microstructures of plasma-sprayed ceramic coatings, while Cr electroplating and HVOF-sprayed cermets chemical and microstructural characteristics have already been described elsewhere [29]. Obvious microstructural differences appear between the various plasmasprayed ceramic coatings. Cr2 O3 markedly exhibits a lamellar microstructure with a prevalence of elongated interlamellar pores; while almost no unmolten particles are found (Fig. 1A).

The overall coating porosity from image analysis is quite low (about 6%): the employment of powder particles with small average grain diameter reduces the average size of pores due to splats stacking faults, gas entrapment and unmolten particles. In Al2 O3 (Fig. 1B), the interlamellar cohesion appears higher. The porosity is still about 6%, due to more rounded pores produced by unmolten particles, splats stacking faults and gas entrapment. In Al2 O3 TiO2 , TiO2 was well melted and partly mixed with alumina, since SEM micrographs highlight lamellar areas containing different amounts of titania. Therefore, intersplat adhesion is better than the former coatings, but many vertical microcracks are present. Unmolten alumina particles can also be found due to the coarser average particle size of this powder; they cause splat-stacking faults. Thus, due to these defects, the porosity of this coating is the highest among the presently tested ceramic coatings (about 9%), notwithstanding the better intersplat adhesion. The XRD qualitative analysis of the starting powders showed that the chromia powder fully consisted in eskolaite phase (Cr2 O3 ), alumina powder in Al2 O3 and aluminatitania powder in -Al2 O3 and anatase, i.e. titanium dioxide low temperature phase. No glass phase was present in any powder. From XRD, the chromia coating consists in eskolaite; the alumina one mainly consists in -Al2 O3

Fig. 1. SEM micrographs of coatings cross-sections. (A and B) Plasma-sprayed Cr2 O3 + NiCoCrAlY bond coat; (C and D) plasma-sprayed Al2 O3 + NiCoCrAlY bond coat; (E and F) plasma-sprayed Al2 O3 13%TiO2 + NiCoCrAlY bond coat.

1302

G. Bolelli et al. / Wear 261 (2006) 12981315

Table 4 Vickers microhardness, indentation fracture toughness, elastic modulus (OliverPharr formula), Ra, Rz for all tested coatings HV1N (GPa) Al2 O3 Al2 O3 TiO2 Cr2 O3 WCCo WCCoCr Hard Cr 11.70 8.18 12.52 11.16 10.58 8.42 1.70 1.29 1.26 0.41 0.46 0.20 KIC (MPa m1/2 ) 2.57 0.65 1.66 0.45 1.67 0.67 4.02 0.51 3.76 0.46 Not measurable (ill-shaped microcracks) [29] Not measurable (coating too thin) Not measurable (coating too thin) E (GPa) 184 141 185 255 257 188 6 7 13 9 16 8 Ra ( m) 5.00 0.82 8.26 1.29 3.07 0.47 3.25 0.28 3.41 0.34 0.518 0.113 2.77 0.24 Same as NiP Ra ( m) ground 0.102 0.105 0.099 0.087 0.080 0.050 0.055 0.047 0.036 0.033 Rz ( m) 18.33 2.70 30.46 4.71 11.60 1.03 12.75 0.66 12.60 1.50 3.30 0.87 15.36 1.31 Same as NiP Rz ( m) ground 2.435 2.967 2.121 0.917 0.775 0.547 0.961 0.348 0.389 0.286

NiP NiP tt

5.52 0.62 9.17 0.09

224 21 184 6

(due to splats quenching), with minor quantities of -Al2 O3 and glassy phase; the aluminatitania consists mainly in -Al2 O3 with some -Al2 O3 , glassy phase (Fig. 2, see black arrow) and a minor amount of rutile. The very low amount of crystalline TiO2 indicates that it mostly dissolves in molten Al2 O3 , as SEM micrographs suggested, contributing to the formation of a low melting point glassy phase, which is probably responsible for the better intersplat cohesion. The hardness, fracture toughness, elastic moduli and roughness parameters of the tested coatings are listed in Table 4. It should be noticed that plasmasprayed ceramic coatings have different roughness: roughness seems to be increasing with increasing powder average particle size. The roughness of plasma-sprayed Cr2 O3 , HVOF-sprayed cermets and electroless NiP plating is quite similar, while electrolytic hard chrome possesses much lower roughness [29]. Cr2 O3 is the hardest of all tested coatings; Al2 O3 has similar hardness to WCCo, while WCCo-Cr is slightly less hard. The microhardness of Al2 O3 is lower than that of bulk alumina (HV = 20.45 GPa [40]), both because of the intrinsically lower hardness of -Al2 O3 than -Al2 O3 and because the indentation response of a plasma-sprayed material is governed not only by the intrinsic hardness of the material, but also by the lamellar microstructure, with splat boundaries giving off under load to facilitate the indenter accommodation [41]. Electroplated chrome is less hard than all thermally sprayed coatings except for Al2 O3 TiO2 . NiP has poor hardness in the as-deposited con-

dition, but is denitely improved after the thermal treatment, due to the two-phase nanocrystalline structure. Al2 O3 possesses the best fracture toughness among plasma-sprayed ceramics, while Al2 O3 TiO2 and Cr2 O3 are less tough. Cermets are much tougher than plasma-sprayed ceramics. The effect of the thermal treatment on electroless Ni is immediately obvious from the huge microhardness increase. Cracked Vickers microindentations are shown in Fig. 3AC for plasma-sprayed ceramics: it can be seen that, for alumina and chromia thermally sprayed coatings, crack preferentially propagate parallel to the substrate interface, along splat boundaries. This is particularly evident for Cr2 O3 , where SEM micrographs had already highlighted a low intersplat cohesion. The abovementioned phenomenon is typical also of HVOF coatings. The aluminatitania coating seems to be the most isotropic among all tested coatings, with similar microcracks propagating both parallel and transverse to the substrate (Fig. 3C). 3.2. Dry particles abrasion resistance The dry sand-steel wheel test results are listed in Table 5. Both wear volume (mm3 ) and wear rate (mm3 /Nm) for increasing number of steel wheel revolutions have been included. While for ceramic coatings the wear rate remains quite constant independently of the number of disk revolutions, for hard chrome plating and HVOF-sprayed cermets there is a clearly decreasing trend, which is better highlighted in Fig. 4. Comparing the observed wear rates for the various coatings, it can be noticed that electrolytic hard chrome and HVOF-sprayed WCCo rank similarly in this test, with WCCoCr (the less tough among tested cermets) performing worse. At a high number of disk revolutions, such coatings behave better than plasma-sprayed ceramics, but at a low number of revolutions, oxide ceramics perform denitely better, with the wear rate of the toughest one, i.e. Al2 O3 , being much lower than that of Al2 O3 TiO2 and Cr2 O3 . The wear scar on plasma-sprayed alumina is shown in Fig. 5, while Fig. 4 compares the wear rate of plasma-sprayed alumina and of HVOF-sprayed WCCo cermet as a function of number of disk revolutions, as a conrmation to the above observations. Wear scar analysis also indicates that plasma-sprayed ceramics (in particular Al2 O3 and Cr2 O3 ), due to their high hardness, do not undergo signicant plastic deformation-related

Fig. 2. X-ray pattern of aluminatitania plasma spray coating. The black arrow indicate the broad band caused by the presence of a glassy phase. The labels are the following: = -Al2 O3 , = -Al2 O3 , = -Al2 O3 , R = rutile (TiO2 ).

G. Bolelli et al. / Wear 261 (2006) 12981315

1303

Fig. 3. SEM micrographs of high-load, cracked Vickers microindentations used for fracture toughness measurement. (A) Plasma-sprayed Cr2 O3 + bond coat; (B) plasma-sprayed Al2 O3 + bond coat; (C) plasma-sprayed Al2 O3 13%TiO2 + bond coat;.

phenomena (such as microcutting and microploughing): their only relevant wear mechanism is brittle fracture, and in particular, cracks mostly propagate through splat boundaries, which are the weakest link in the material, as former indentation tests clearly showed. 3.3. Pin-on-disk wear test Hertzian maximum and average contact pressures, maximum sub-supercial shear stress and maximum shear stress depth are all reported in Table 6; they have been computed from analytical Hertzs formulae, using the experimentally evaluated elastic moduli in Table 4 for the samples and assuming E = 210 GPa, = 0.3 for 100Cr6 steel, E = 380 GPa, = 0.23 for sintered alumina. Clearly, such stresses are valid in a static contact condition,

but do not take into account additional stresses due to friction, which make the contact more severe. It must be considered that hertzian stresses and pressures are only valid at the beginning of the test; in fact, as the test progresses, the pin gets progressively worn, the actual contact area increases and the contact pressure consequently decreases. It can be noticed that differences between contact pressures and contact stresses in the various coatings arise due to the different elastic moduli; however, in the same test, they always remain the same order of magnitude, the highest shear stress differences being about 25% between WCCoCr and Al2 O3 TiO2 . Most importantly, the maximum shear stresses always occur within the coating, without any signicant involvement of the substrate and the coating-substrate interface. Thus, even though tested coatings have different thicknesses, pin-on-disk test results can be compared. The pin-on-

Table 5 wear volumes (mm3 ) and wear rates (103 mm3/Nm) recorded in dry sand-steel wheel abrasion test Test/sample 7 revolutions Wear volume Wear rate 40 revolutions Wear volume Wear rate 60 revolutions Wear volume Wear rate 80 revolutions Wear volume Wear rate Average wear rate Al2 O3 2.03 0.24 11.47 1.38 10.46 1.65 10.37 1.64 15.26 1.96 10.09 1.29 20.08 2.16 9.96 1.07 10.14 1.29 Al2 O3 TiO2 Not measured because the wear rate stays constant 14.97 0.88 14.85 0.87 21.82 0.96 14.43 0.64 24.24 1.47 12.02 0.73 13.73 1.47 Cr2 O3 Not measured because the wear rate stays constant 16.88 1.87 16.74 1.85 22.15 2.50 14.65 1.65 27.26 3.18 13.52 1.58 14.93 2.12 WC17%Co 4.27 0.01 24.19 0.01 8.44 0.02 8.37 0.02 9.61 0.01 6.36 0.01 10.82 0.27 5.37 0.14 Not calculated WC10%Co4%Cr 6.79 0.07 38.50 0.38 13.48 0.04 13.34 0.04 15.30 0.50 10.12 0.33 16.70 0.35 8.28 0.17 Not calculated Hard Cr 3.81 0.15 21.63 0.82 7.52 0.22 7.46 0.22 8.72 0.17 5.76 0.11 10.20 0.13 5.06 0.07 Not calculated

1304

G. Bolelli et al. / Wear 261 (2006) 12981315

Fig. 4. Al2 O3 and WCCo wear rates in dry sand-steel wheel tests as a function of number of disks revolutions.

Fig. 5. SEM micrograph of dry sand-steel wheel wear scar on plasma-sprayed Al2 O3 . Black arrows indicate some particularly evident splats, proof of the brittle detachment of nearby splats.

disk results for plasma-sprayed coatings are shown in Fig. 6A (sample wear rate), Fig. 6B (pin wear rate) and Fig. 6C (average friction coefcients). Material build-up on the coatings surface is recorded on all samples when tested against 100Cr6 due to steel debris transfer from pin to sample, while the pin undergoes signicant material loss. The occurrence of material build-up is indicated in Fig. 6A as a negative value of the wear rate. The Cr2 O3 coating causes much less wear on the steel pin (subsequently undergoing much less material transfer from the pin) and displays lower friction coefcients. The wear rate of the steel pin

against alumina and aluminatitania is 210 times higher than that against chromia in all cases (except for Al2 O3 TiO2 in test 7); the wear rate against Cr2 O3 never exceeds 105 mm3 /Nm. The alumina pin undergoes lower wear rates (106 mm3 /Nm) when sliding against Cr2 O3 as well. Overall, Cr2 O3 coating outperforms other plasma-sprayed ceramics in the pin-on-disk test, concerning both wear rates (undergone and inicted on pins) and friction coefcients. The SEM micrographs of wear scars of plasma-sprayed coatings tested (test 7) against 100Cr6 (Fig. 7A for Al2 O, Fig. 7B for Cr2 O3 ) conrm that, during

Table 6 Hertzian maximum and average contact pressures, maximum sub-supercial shear stress, and depth of the point of maximum shear stress, for all pin-on-disk test congurations Maximum hertzian contact pressure (MPa) Al2 O3 Test 26: 716.9 Test 48: 1225.8 Test 15: 620.8 Test 37: 1061.4 Test 26: 632.9 Test 48: 1082.2 Test 15: 560.5 Test 37: 958.4 Test 26: 718.6 Test 48: 1228.8 Test 15: 622.0 Test 37: 1063.6 Test 48: 1430.4 Test 37: 1198.5 Test 48: 1434.8 Test 37: 1201.4 Test 48: 1259.8 Test 37: 1085.0 Test 48: 1357.6 Test 37: 1151.1 Test 48: 1247.9 Test 37: 1076.8 Average hertzian contact pressure (MPa) Test 26: 477.9 Test 48: 817.2 Test 15: 413.4 Test 37: 707.7 Test 26: 421.9 Test 48: 721.5 Test 15: 373.7 Test 37: 639.0 Test 26: 479.1 Test 48: 819.2 Test 15: 414.7 Test 37: 709.1 Test 48: 953.6 Test 37: 799.0 Test 48: 956.5 Test 37: 800.9 Test 48: 839.9 Test 37: 723.4 Test 48: 905.1 Test 37: 767.4 Test 48: 831.9 Test 37: 717.9 Maximum hertzian shear stress (MPa) Test 26: 222.2 Test 48: 380.0 Test 15: 192.4 Test 37: 329.1 Test 26: 196.2 Test 48: 335.5 Test 15: 173.8 Test 37: 297.1 Test 26: 222.8 Test 48: 380.9 Test 15: 192.8 Test 37: 329.7 Test 48: 443.4 Test 37: 371.5 Test 48: 444.8 Test 37: 372.4 Test 48: 390.5 Test 37: 336.4 Test 48: 420.9 Test 37: 356.8 Test 48: 386.8 Test 37: 333.8 Maximum shear stress depth ( m) Test 26: 12.4 Test 48: 21.2 Test 15: 13.3 Test 37: 22.8 Test 26: 13.2 Test 48: 22.5 Test 15: 14.0 Test 37: 24.0 Test 26: 12.4 Test 48: 21.2 Test 15: 13.3 Test 37: 22.7 Test 48: 19.6 Test 37: 21.4 Test 48: 19.6 Test 37: 21.4 Test 48: 20.9 Test 37: 22.5 Test 48: 20.1 Test 37: 21.9 Test 48: 21.0 Test 37: 22.6

Al2 O3 TiO2

Cr2 O3

WCCo WCCoCr Cr NiP NiP tt

G. Bolelli et al. / Wear 261 (2006) 12981315

1305

Fig. 6. Pin-on-disk test results for plasma-sprayed coatings. (A) Coatings wear rates. Negative wear rates indicate material build-up on coating surface due to wear debris sticking; (B) pin wear rates; (C) friction coefcients.

the test, some steel debris is transferred to the coatings. When ceramic coatings are tested against alumina pins, a tribolm is formed (Fig. 8AC), consisting in plastically deformed debris from both sample and pin, and also in plastically deformed ceramic splats (Fig. 9AD). On Al2 O3 and Al2 O3 TiO2 coatings (Fig. 8A and B), besides plastically deformed splats, lot of plastically deformed debris appears in the surface lm, which seems to possess low compactness (Fig. 9A and B). On Cr2 O3 , the surface lm is much more compact and smooth in all cases (Fig. 8C): it mostly consists in plastically deformed debris and splats (Fig. 9C and D), with some alumina debris, which were also plastically deformed and strongly embedded in this lm. Optical micrographs of pin wear scars are in Fig. 10. They high-

light that the steel sample mostly suffered two-body grooving (ploughing and cutting) but also some adhesive wear. Alumina, instead, mostly undergoes brittle fracture; besides, some plastically deformed wear debris adheres to the pin, forming an irregular surface lm, which undergoes grooving. Fig. 11A and B represent three-dimensional axonometric projections of wear scars on Al2 O3 and Cr2 O3 after test no. 8, reproduced from the optical prolometry results. While Al2 O3 has undergone signicant material loss, with pin penetration inside the coating, the tribolm on Cr2 O3 is smoother than the original surface and is placed slightly above the mean line of the unworn surface. Table 7 sums up the wear rates of all tested samples in the pin on disk wear test. HVOF-sprayed coatings exhibit an

Fig. 7. SEM micrographs of wear scars on plasma-sprayed Al2 O3 (A) and Cr2 O3 (B) after test no. 7, indicating transferring of steel debris on the coating surface.

1306

G. Bolelli et al. / Wear 261 (2006) 12981315

Fig. 8. SEM micrographs of wear scars on plasma-sprayed Al2 O3 (A), Al2 O3 13%TiO2 (B) and Cr2 O3 (C) after test no. 8, showing the formation of tribolms.

excellent sliding wear resistance: they never undergo a measurable volume loss. Sometimes, material build-up is recorded from optical prolometry. Their performance is therefore comparable to Cr2 O3 . Electrolytic hard chrome displays unfavourable wear and friction characteristics when compared to plasma-sprayed Cr2 O3 and HVOF-sprayed cermets: when tested against sintered alumina, it undergoes a signicant wear rate (but lower than Al2 O3 and Al2 O3 TiO2 ). Remarkably, thermally treated

NiP undergoes a negligible wear rate in all tests. Without thermal treatment, it undergoes very high wear rates both against alumina and against 100Cr6, being less hard than steel. However, its wear rate becomes negligible in test no. 8. The reasons for this behaviour will be discussed in the following section. Fig. 12A and B respectively compare pin wear rates and average friction coefcients for plasma-sprayed ceramic coatings, HVOF-sprayed cermets, Cr electroplating and NiP electroless

Fig. 9. Details of wear scars on plasma-sprayed ceramic coatings after testing against alumina pins. (A) Plasma-sprayed Al2 O3 , test no. 8; (B) plasma-sprayed Al2 O3 13%TiO2 , test no. 6; white arrows indicate plastically deformed splats; (C and D) plasma-sprayed Cr2 O3 , test no. 8, with the appearance of undeformed splats under the tribolm.

G. Bolelli et al. / Wear 261 (2006) 12981315

1307

Fig. 10. Optical micrographs of pin wear scars. (A) Steel pin after test no. 3 against Cr2 O3 ; (B) alumina pin after test no. 4 against Cr2 O3 . Table 7 Sample wear rates in mm3 /Nm after pin on disk test on as-deposited (no grinding) coatings Test 1 Cr2 O3 Al2 O3 Al2 O3 TiO2 Cr WCCo WCCoCr Ni Ni-tt 1.33 103 6.52 103 2.07 104 Test 2 4.98 104 1.02 103 0 1.35 104 Test 3 3.25 105 3.29 104 0 6.56 104 9.86 105 7.14 104 0 6.46 105 Test 4 4.91 105 1.95 104 3.06 104 4.11 105 4.59 105 2.96 105 8.77 104 2.17 105 Test 5 5.64 104 1.42 104 1.33 103 Test 6 2.48 104 1.12 103 6.02 104 0 Test 7 5.79 105 3.12 104 2.36 104 0 1.45 104 1.49 104 6.46 104 0 Test 8 5.12 105 5.80 104 4.72 104 1.90 104 4.58 105 5.18 105 0 0

Note: Negative values indicate material build-up on the sample.

platings. Alumina pins wear rates are higher by almost one order of magnitude when tested against cermets than when tested against plasma-sprayed Cr2 O3 . Electrolytic hard chrome also causes higher friction coefcients and higher pin wear rates than Cr2 O3 when tested against alumina pins (especially with the lower sliding speed), but lower than that caused by cermets. In the tests against 100Cr6 pins, electrolytic hard chrome displays higher friction coefcient than plasma-sprayed chromia and HVOF-sprayed cermets. Reversing the results for alumina pins, 100Cr6 pins undergo one order of magnitude lower wear rates when slid against cermets than when slid against Cr2 O3 . When sliding against alumina, the cermet coatings never undergo signicant material removal, thanks to the formation of a very compact tribolm, located on the higher asperities of the surface. EHC is not hard enough to prevent ploughing and cutting by the alumina pin, though it is tough enough to prevent material removal by brittle cracks propagation from pre-existing cracks. As far as thermally treated NiP electroless plating are concerned, after a short running-in period in which the domes caps are slightly worn (due to very high localized contact pressures),

a lm is formed where an oxidized NiP compound, possessing low chemical afnity to the counterbody and (probably) high hardness, lls the valleys between adjacent domes. Wear scars of NiP electroless platings after tests 4 and 8, respectively, show denite morphological differences, which will be hereafter discussed and explained. It must be noticed that the sample wear scar after the seemingly anomalous test no. 8 is almost identical to that of thermally treated NiP, with slightly worn domes caps and NiPO compounds lling roughness valleys. Alumina pins undergo both two-body grooving and brittle fracture against HVOF-sprayed WCCo and thermally treated electroless Ni plating respectively, with the latter forming a much smoother wear scar on the pin. The effects of polishing on tribological performance of coatings has been the object of a specic investigation, comparing the results of tests 7 and 8 performed both on as-deposited and on ground plasma-sprayed ceramics, HVOFsprayed WCCo and electrolytic hard chrome. The recorded samples and pins wear rates are listed in Fig. 13A, friction coefcients are indicated in Fig. 13B. No wear rate is indicated for samples tested against 100Cr6, because only material build-up

Fig. 11. Three-dimensional axonometric representations of wear scars on plasma-sprayed Al2 O3 (A) and Cr2 O3 (B) after test no. 8.

1308

G. Bolelli et al. / Wear 261 (2006) 12981315

Fig. 12. Pin-on-disk test results for all tested coatings. (A) Pin wear rates; (B) friction coefcients.

is found in this case. It must also be noticed that wear rates for ground Cr2 O3 and WCCo tested against alumina have been indicated, while a limited material build-up was recorded on as-sprayed coatings. The wear rates for WCCo and Cr2 O3 , however, are more than two orders of magnitude lower than those for the other coatings; thus, their wear performance is still far superior to other tested materials. Such very low wear rates could not have been measured on rough samples, but surface grinding has made it possible to determine them accurately. Thus, it can be assumed that the overall wear amount undergone by these samples has not actually increased after grinding: like as-sprayed coatings, ground ones do not show a real wear scar with signicant material removal from the sample surface, but only minor surface morphological changes. Thus, samples wear rates are not signicantly changed in any case, while some differences emerge in pin wear rates and friction coefcients. Pin wear rates decrease in all cases (with the exception of 100Cr6 pin against Al2 O3 TiO2 ). Considering alumina pins, Cr2 O3 remains the ceramic coating inicting less wear to the counterpart, electroplated chromium still causes higher wear (comparable to the other ceramic coatings), but WCCo now causes much less wear than hard chrome and plasma-sprayed Al2 O3 and Al2 O3 TiO2 , approaching the performance of Cr2 O3 . Consid-

Fig. 13. Comparison between pin-on-disk tests no. 7 and no. 8 for as-deposited and ground coatings. (A) Sample wear rates in test no. 8 (only positive wear rates indicating effective material removal are indicated) and pin wear rates in test no. 7 (steel pin) and test no. 8 (alumina pin); (B) friction coefcients.

ering 100Cr6 steel pins, the wear rates that all ceramic coatings and electrolytic hard chrome inict become very similar (about 1.3 105 mm3 /Nm), with WCCo still causing the least wear on the steel pin. The friction coefcients against 100Cr6 pins decrease for plasma-sprayed ceramics (which display similar values after polishing) but increase for electrolytic hard chrome and HVOF-sprayed WCCo. Friction coefcients against alumina pins decrease for all coatings but Al2 O3 and Al2 O3 TiO2 (which display slightly increased friction) and rankings between coatings remain similar, with Cr2 O3 causing the lowest friction coefcient against alumina, followed by HVOF-sprayed WCCo, by hard chrome, and nally by Al2 O3 and Al2 O3 TiO2 . Fig. 14AD show optical micrographs of some worn alumina and steel pins. The steel pin undergoes two-body abrasive wear (cutting and ploughing) when tested against chromia (Fig. 14A); instead, a lm of transferred material is formed on the pin tested

G. Bolelli et al. / Wear 261 (2006) 12981315

1309

Fig. 14. Optical micrographs of pin wear scars after testing against ground coatings. (A) Steel pin after test no. 7 against ground Cr2 O3 ; (B) steel pin after test no. 7 against ground WC17%Co; (C) steel pin after test no. 7 against ground electrolytic hard chrome; (D) alumina pin against test no. 8 against ground Cr2 O3 .

against WCCo (Fig. 14B): it is likely that this lm consists in plastically deformed (and probably also partly oxidized) metallic debris from the pin itself and from the soft Co matrix of the cermet. The pin suffers both abrasive and adhesive wear in tests against electrolytic hard chrome (Fig. 14C). The wear scar on the alumina pin tested against Cr2 O3 (Fig. 14D) and WCCo is very smooth, with some grain boundaries becoming very evident. Prolometry on the wear scar on polished Cr2 O3 and WCCo samples after test no. 8 allowed to notice that the surface roughness of Cr2 O3 has decreased because some asperities have been worn out and pores (which have been opened be surface polishing) have been lled with wear debris. The surface roughness of WCCo, instead, has increased, because some metal matrix has been abraded by alumina pin asperities, leaving the carbide grains protruding out of the surface. 4. Discussion 4.1. Microstructure and micromechanical properties Among plasma-sprayed coatings, Cr2 O3 seems to possess the lowest intersplat cohesion, as emerging from SEM micrographs (Fig. 1A and B) and from fracture toughness measurement. However, from the same SEM micrographs, it seems that unmolten particles are not the reason for the low intersplat cohesion of this coating; in fact, all splats appear well spread, indicating good in-ight melting. The as-sprayed coating roughness conrms a very good molten droplets attening for this coating. It can be argued that the well-spread splats may have too low roughness to allow proper mechanical adhesion of new-coming

splats. This phenomenon has already been observed with HVOFsprayed alumina splats [42], which spread very well due to their high in-ight velocity and thus possess low mechanical interlocking. Furthermore, the quite high melting point of Cr2 O3 (2330 C [40]) causes molten splats to solidify very quickly upon impact and does not allow underlying splats to reach a sufciently high temperature to activate chemical intersplat bonding. In fact, it is noticeable, from the vast literature on plasma spraying of ceramics (for instance, papers dealing with plasma-sprayed PSZ [4347] and Al2 O3 TiO2 [25,48] can be compared) that the more high-melting-point the ceramic powder is, the more difcult it is to achieve a dense, low porosity coating. In the alumina coating, where intersplat cohesion is better, overall fracture toughness is higher (Table 4 and Fig. 3B) because crack propagation along splat boundaries is more difcult. The low melting point of titania in Al2 O3 TiO2 coating favours better particle melting, causing opposing effects. On the one hand, it improves intersplat adhesion, as Fig. 1C indicates, conrming many literature work [25,48]. Probably it also promotes chemical intersplat bonding as well. Thus, the abovementioned phenomena lessen coating anisotropy. On the other hand, however, the addition of titania allows the formation a brittle glassy phase inside the coating. Furthermore, the better intersplat adhesion also causes more transverse microcracking due to tensile quenching stresses. Besides, the coarser average powder particle size causes the appearance of a signicant number of unmolten particles. This is why signicant cracks both parallel and transverse to the substrate are formed even at a lower load than other plasma-sprayed coatings (5 N instead of 10 N). Overall, the result is a less tough and less hard coating than

1310

G. Bolelli et al. / Wear 261 (2006) 12981315

plasma sprayed alumina alone. It is very important to notice that better as-sprayed coating surface nishing can be achieved using ner spray powders; in fact, even though chromia is the most high melting point among tested ceramics (thus, in theory, the most difcult to melt), it is the one for which the best surface nishing has been obtained in the as-sprayed state. Having a lower as-sprayed surface roughness is very important for technological applications, because it reduces the number of post-deposition mechanical treatments necessary. HVOF-sprayed cermets are tougher than plasma-sprayed ceramics thanks to their composite nature with hard particles in a tough metal matrix and thanks to a better cohesion, allowed by much higher particles speed in the supersonic HVOF ame. Nonetheless, the coatings are still denitely anisotropic, with cracks mostly propagating parallel to the substrate. These micromechanical features shall obviously play an important role in explaining the results of tribological tests. 4.2. Dry particles abrasion resistance Former research has shown that, in such kinds of tests, the decreasing wear rate of coatings containing metallic phases is caused by abrasive particles which are progressively embedded in metals or cermets due to the ductility of the metallic phase [29,49]. This phenomenon has not limited to this specic test condition, but it has been also documented under different experimental conguration (microscale abrasion) [50]. In this way, the wear rate is decreased with increasing sliding distance because embedded particles alter the characteristics of the sample surface, protecting it. Thus, wear rates for ceramic coatings can be calculated as the average of measurements performed at different numbers of disk revolutions, but this is not possible for metals and cermets. The wear mechanism for hard chrome plating under the dry sand-steel wheel test is mixed three-body grooving and two-body rolling (following literature denitions discussed for instance in [51]), with indenting abrasive particles getting progressively embedded in the coating. The coating thus undergoes microcutting, microploughing and also microindentation. It can be seen that the chrome plating is not hard enough to prevent hard alumina grains to wear the surface by grooving phenomena (ploughing, cutting), but withstands indentation without cracking thanks to its sufcient toughness. HVOF-sprayed cermets undergo two and three-body wear as well, as already discussed in [29]. Ceramic coatings do not undergo ploughing and cutting, because of the very high intrinsic hardness of the material, comparable to the one of the abrasive corundum grains; instead, they undergo wear due to brittle splats removal. Therefore, their abrasive wear resistance overcomes that of hard chrome plating and HVOF-sprayed cermets, for a low number of disk revolutions, while at high number of disk revolutions the embedded particles protect the metallic and cermet coatings from further grooving or indenting phenomena. So, in the latter conditions, cermets and hard chrome plating wear rate is better than ceramics. Since the wear rate for ceramic coatings remain quite constant and the wear mechanism is not altered increasing the test duration, the steel wheel test can be considered signicant for them. Plasma-sprayed ceramic coatings, therefore, seem a technically

appropriate choice when dry particles abrasion is the main wear mechanism involved. The best performance for ceramic coatings is shown by the toughest one. It would therefore be appropriate to elucidate the relationship between fracture toughness, Vickers microhardness and abrasive wear resistance, to quantitatively ascertain the effect of microstructural characteristics on this property, in view of an optimization for practical application. An anomaly is immediately apparent when observing that Al2 O3 13%TiO2 is more abrasion resistant than Cr2 O3 , which is harder and possesses similar toughness. Two facts must be considered to explain this seeming inconsistency: the main wear mechanism is splats detachment along splats boundaries; thus, a quantity directly related to intersplat cohesion should be adopted; in fracture toughness tests, long cracks in the direction parallel to the substrate (i.e. along splat boundaries) appeared for Cr2 O3 , while similar cracks were developed in al directions for aluminatitania. Thus, fracture toughness by itself might not be the best parameter to explain abrasive wear behaviour; a parameter taking into account the toughness in the direction parallel to the substrate and the coating anisotropy must be introduced. An empirical but simple parameter can be computed by using only the half-diagonal and average crack length in the direction parallel to the substrate in the EvansCharles formula, thus obtaining an estimation of the crack resistance in this direction. This parameter will be hereafter referred to as KIC-L (MPa m1/2 ). The KIC-L values and the KIC-L /KIC ratio are listed in Table 8. The KIC-L /KIC ratio will be employed in a modied version of the EvansMarshall formula for abrasive wear rate prediction [52]: w=a (E/H)P 1/8 KIC H 5/8
1/2

(4.1)

where a is the material independent constant, w = wear rate (mm3 /Nm), E the elastic modulus (GPa), H the Vickers microhardness (GPa), P the normal load (N), and KIC is the fracture toughness (MPa m1/2 ). The formula will be modied by multiplying the denominator by the KIC-L /KIC ratio. Since the normal load has not been changed, its effect cannot be accounted for and will be considered as a part of the coefcient a. Since the formula has been changed, it is likely that the exponents value shall have to be changed as well. The resulting expression is as follows: w = a1 (E/H)
a2 KIC (KIC-L /KIC )a3 H a4

(4.2)

The coefcient a1 and the a2 , a3 , a4 exponents have been evaluated by non-linear least squares tting of experimental data: a1 = 0.0121, a2 = 0.1290, a3 = 0.7355, a4 = 1.0290.
Table 8 KIC-L values and KIC-L /KIC ratios for the plasma-sprayed ceramic coatings Coating KIC-L KIC-L /KIC (MPa m1/2 ) Al2 O3 1.76 0.88 0.68 0.34 Al2 O3 TiO2 3.30 1.23 1.28 0.48 Cr2 O3 0.85 0.40 0.51 0.33

G. Bolelli et al. / Wear 261 (2006) 12981315

1311

Fig. 15. Dry sand-steel wheel test results for plasma-sprayed ceramic coatings tted with Eqs. (4.1) and (4.2). (A) Wear rates plotted 1/2 against (E/H)P 1/8 /(KIC H 5/8 ); (B) wear rates plotted against (E/H) a2 (KIC (KIC-L /KIC )a3 H a4 ).

Fig. 15A shows that the EvansMarshall equation fails in predicting the abrasive wear rates of the various coatings, while the new equation (Fig. 15B) ts the experimental data. It should be noticed that, in the current work, only three points have been employed to t Eq. (4.2), while, for a proper statistical analysis, a greater number of points should have been employed. The t results, therefore, should not be considered as being valid in general. The above considerations only seek to indicate a new way for the microstructural interpretation of abrasive wear test results of plasma-sprayed coatings. Although there exist doubts concerning the effective usefulness of such predictive theoretical or semiempirical formulae for thermally sprayed coatings [53,54], they can be a useful reference for a simple estimation of the performance of a coating, and a basis for a future a-priori process design, once the correlations between operating parameters and micromechanical properties shall be fully elucidated by further fundamental research. 4.3. Pin on disk wear test It is a basic principle that, in a tribological coupling between two different surfaces, the harder one wears the softer one: since all tested coatings but NiP plating without treatment are harder than 100Cr6, they never undergo a wear loss except for NiP

without thermal treatment. The only way by which material could be removed from samples surface would be by crack formation and propagation by cyclic contact fatigue; however, it appears that contact stresses have not been high enough to cause such phenomenon. Tribological phenomena in couplings between the steel pin and tested coatings seem to be inuenced by the coating roughness. In many cases, the steel pin forms a lm of transferred material on the coatings, thus, after a runningin period, a steelsteel contact governs tribological phenomena. The morphology of the transferred lm is inuenced by coating surface nishing: considering as-sprayed ceramic coatings, the lm is thinner and smoother on the Cr2 O3 one, which possess the lowest surface roughness, while it is coarser and thicker on the others. This explains the lower friction coefcient recorded on the as-sprayed chromia coating. When ceramic coatings are ground, they all cause a similar friction coefcient and pin wear against the 100Cr6 pin, because they all form a quite smooth and thin lm of transferred material. On the electrolytic hard chrome plating, no signicant transferred lm is formed, but the friction coefcient is the highest among tested coatings: in this case, chemical afnity between the surfaces must be considered. While ceramic coatings have no chemical afnity for the 100Cr6 pin, thus adding no further source of friction and wear, a denite chemical afnity between the two metallic surfaces of electroplated chromium and 100Cr6 certainly exists. This is conrmed by optical micrographs indicating adhesive wear taking place on the steel pin in this case. Therefore, when the chrome plating is polished, the friction coefcient slightly increases, because the overall extent of the actual contact area is increased. The same considerations are valid for as-deposited NiP platings as well; after the thermal treatment, instead, the appearance of the Ni3 P phase and the formation of oxidized phases on the coating surface denitely lowers the chemical afnity, reducing the friction coefcient. A very peculiar situation occurs for cermet coatings: in the as-sprayed condition, the contact is localized on coating surface asperities, so that the pin is initially worn by few protruding carbide particles and the steel debris form a transfer lm on the coating. After polishing, the contact area becomes much larger and also involves the soft Co matrix. In this situation, the chemical afnity between the soft Co matrix and the steel pin becomes relevant, a lm of transferred material is formed on the pin itself and the friction coefcient increases, although it remains lower than that for hard chrome and as-deposited NiP, both because of transfer material oxidation and because a large part of the cermet surface consists of ceramic WC particles possessing low chemical afnity for steel. The friction coefcient and pin wear rate decrease with increasing contact load for almost all coatings and with increasing sliding speed as well, for ceramic coatings. Considering as-sprayed ceramic and cermet coatings, where a transferred steel lm is formed, under a higher contact load the transferred lm on the sample surface is more rapidly built up, reducing wear caused by the hard coating asperities. Furthermore, under a higher normal load, a higher local temperature is reached in the contact point; thus, the steel surface is more rapidly oxidized, reducing the friction coefcient. The latter event also occurs for increased sliding speed. For electrolytic hard chrome, where no signicant transferred

1312

G. Bolelli et al. / Wear 261 (2006) 12981315

steel lm appears, it is likely that the increased contact point temperature favours metallic surfaces oxidation, slightly reducing chemical afnity troubles. Sintered alumina is harder than all tested ceramic coatings; therefore, it is able to inict wear on them and, at the same time, it undergoes much less wear than 100Cr6 (wear rates 105 mm3 /Nm), Fig. 10B. The pin undergoes grooving phenomena (ploughing and cutting) which are probably due to the formation of a thin, softer surface lm consisting in wear debris, and also undergoes a limited extent of brittle cracking and grain pull-out, with rougher wear scars for higher recorded wear rates. Comparing the present wear rates and wear mechanisms with literature work on wear of sintered ceramics [3032], the present alumina pin wear seems in a border-line situation between mild and severe wear. The pin wear rate seems to be increasing both with normal load, which favours brittle cracking, and with sliding speed, which increases ash temperature in the contact point favouring thermal shock cracking [30], although a few exceptions exist in the present experimental data set. Obviously, since brittle fracture is involved in alumina pin wear, reducing the contact pressure on surface asperities by sample grinding lowers wear alumina pin wear rates and also lowers friction coefcients. This is particularly obvious for the WCCo cermet: the wear rate it inicts on the alumina pin decreases by more than two orders of magnitude after grinding. The measured wear rate (<107 mm3 /Nm) and the smooth pin wear scar, with no brittle cracking, clearly indicates the occurrence of a mild pin wear regime (according to [3032]) in tests against ground WCCo and Cr2 O3 . Friction coefcient and pin wear are quite high for electrolytic hard chrome and for Al2 O3 and Al2 O3 TiO2 coatings: for these latter, chemical afnity is likely to be the dominating effect. For hard chrome, some wear debris from coating material is transferred to the alumina pin, so that a metalmetal contact arises, increasing the friction coefcient. Tests against alumina pins also allow to evaluate the wear resistance of the various coatings, since alumina is hard enough to inict wear on most of them. It appears that surface tribolms formation is an important tribological phenomenon in most cases: these tribolms consist in plastically deformed wear debris, plastically deformed coating material or in a surface chemical alteration of the coating. If a tribolm is not formed, the coating cannot oppose to continuous material removal by the harder counterpart asperities, as for electrolytic hard chrome and as-deposited NiP. The tribolm formation, however, is only benecial if the tribolm itself possesses adequate cohesion. Let us examine the case of as-sprayed ceramic coatings. Their tribolm consists in plastically deformed splats and wear debris: plastic deformation is caused by high local contact pressures and by high local contact temperatures, which are due to heat generation by friction and to poor thermal conductivity of ceramic materials. A signicant compressive stresses distribution appears just below the surface at contact points because of high contact pressures: it is known that, under high idrostatic stress, ceramics can plastically deform at the microscale. Besides, high local temperatures lessen the asperities hardness, favouring plastic deformation. Tribological studies on both plasma-sprayed [23,24,55] and massive [3034] ceram-

ics conrm the occurrence of microscale plastic deformation in ceramics: material sliding at discrete shear faults (for instance, grain boundary sliding) is thought to be the cause of such phenomenon. However, the wear scar appearance in massive ceramics, as the above mentioned literature works report and the present observations of alumina pin wear scars conrm, is much different from the one in plasma-sprayed coatings: in massive ceramics, the plastically deformed lm is very thin and compact, with few debris embedded in grain boundaries. In plasma-sprayed coatings, instead, a thicker lm appears, with a higher amount of debris on all the contact surface: this is probably due to material detachment from surface asperities, because of the quite high initial surface roughness. Besides, the splats themselves seem to undergo extensive plastic deformation: it is known that, due to rapid cooling, splats consist in very small crystals (1 m) [16]; thus, a high overall grain boundary surface provides a great amount of shear faults for plastic deformation; moreover, splat boundaries themselves may constitute favourable areas for plastic slipping [23,24]. With this assumption, plastic deformation without sub-surface cracking is more favoured in the Cr2 O3 coating, where splat cohesion is lower. Thus, surface splats on the chromia coating are quite free to deform plastically without sub-surface cracking, while, on other coatings, splats are less free to deform plastically and are more easily cracked due to elastic stresses accumulation. Thus, on Al2 O3 and Al2 O3 TiO2 , the lm has a quite loose structure. During the test, it is continuously removed and reformed and does not protect the coating, accounting for the higher wear rate. On plasma-sprayed Cr2 O3 , the compact lm effectively protects the surface from damage; in fact, SEM micrographs (Fig. 9C and D) indicate that, below this lm, there are unaltered splats. Furthermore, the lm smoothness may partly account for the lower friction coefcient recorded in this case. Another reason for the lower friction coefcient might be the appearance of CrO2 and CrO3 on the very surface of the lm, reducing the friction coefcient [55]. Furthermore, the higher hardness of the Cr2 O3 coating implies the compact tribolm, once formed, cannot be easily abraded by pin surface asperities or loose debris. On cermets, the occurrence of a tribolm has already been documented in literature [56,29], and, as explained in [29], its formation and its compactness are the effect of the exceptional ductility of the Co-based metal matrix. The metal matrix, in fact, is plastically deformed around contact points because of high contact stresses, and part of the soft metal is abraded by asperities on the counterbody. The plastically deformed metal strongly holds the small carbide particles, some of which have been cracked but, nonetheless, are still embedded. Such particles slightly protrude out of the metal matrix and, being very small and close to each other, prevent further abrasion of the metal matrix. The detached metal debris (together with some very small WC fragment) mix with the alumina debris and acts as bonding phase, allowing debris sticking onto the cermet surface. Therefore, lm areas consisting in alumina particles and WC fragments held by metal matrix are found next to the protruding particles. This lm prevents any further wear on the coating, in fact, the contact is localized on WC particles, which are harder than the alumina counterbody. The hardness of these carbide particles,

G. Bolelli et al. / Wear 261 (2006) 12981315

1313

on the other hand, may cause some wear on the pin (Fig. 12A). As far as thermally treated NiP is concerned, the tribolm is formed by chemical alteration of the coating surface: once the domes caps are smoothed by high contact pressures, the contact point temperature is high enough to form oxidized NiPO compounds which denitely lessen the friction coefcient and contribute (together with the high hardness and toughness of the nanostructured coating) to arrest any further coating wear. The response from the NiP electroless plating in test 8, where the pin and sample wear rate and friction coefcient resemble thermally treated NiP, deserve a special comment. The contact temperature must have been high enough to cause nanocrystallization along the contact area, so that in the contact area, the coating behaves almost exactly like the thermally treated one. It is interesting to observe how an improves surface nishing alters the wear behaviour. On Cr2 O3 , the higher actual contact area reduces initial contact stresses around surface asperities, so that less debris appears. Thus, the lm only consists in plastically deformed splats, with few debris lling open pores. As a result, a very smooth wear scar appears, and no material build-up is recorded, but only a very low wear rate, as shown in Section 3. On WCCo, the higher actual contact area prevents local plastic deformation of the metal matrix around surface asperities, as it happened in as-sprayed coating. Besides, on the ground surface, carbide particles are already slightly protruding, because the soft Co matrix is much more easily ground and removed. Thus, the contact is already located on the very hard carbide particles. Thus, no tribolm formation occurs because the favourable condition of carbides slightly protruding out of the metal matrix has already been accomplished by the grinding operation: the ground surface itself acts as the tribolm. The only noticeable wear phenomenon is the removal of a small amount of Co matrix by the highest counterpart asperities: once the carbide particles are protruding out enough and the counterpart surface asperities are worn away, no further wear occurs on the coating. The smoothness of the wear scars, little amount of wear debris and no material build-up for ground Cr2 O3 and WCCo coatings is consistent with the onset of a mild wear regime for the alumina pin. This accounts for the absence of material build-up on the surface and for the very low measured wear rate. Some literature work [57,58] indicates that wear due to brittle crack propagation could occur: in this case, the contact pressure was not high enough. Besides, crack propagation was reported to be caused by metal matrix embrittlement due to carbides dissolution: in this case, the optimal choice of spraying parameters prevented an excessive extent of carbides dissolution, so that the coating is tough enough to prevent crack propagation under reasonable contact loads. It is practically unusual to nd application where hertzian contact pressures are allowed to exceed 1 GPa [59]. 5. Conclusions Three plasma-sprayed oxide ceramic coatings, namely Cr2 O3 , Al2 O3 , Al2 O3 13%TiO2 , have been characterized in terms of microstructure and micromechanical properties. Their dry particles abrasion resistance (dry sand-steel wheel test) and pin-on-disk wear test response with 100Cr6 steel and

sintered alumina counterbodies (sample wear rate, pin wear rate, friction coefcient) have been compared to the behaviour of two HVOF-sprayed cermet coatings (WC17%Co and WC10%Co4%Cr), electrolytic hard chrome, untreated and thermally treated electroless NiP coatings. The experimental results lead to the following conclusions: Cr2 O3 coating is the hardest and most anisotropic among plasma-sprayed ceramics, due to low interlamellar cohesion; Al2 O3 13%TiO2 is the most isotropic but also the less hard and less tough one, due to the formation of an aluminatitania glassy phase which favours intersplat adhesion but turns out to be quite brittle. WC17%Co and WC10%Co4%Cr are also anisotropic, but, thanks to a ductile metal matrix and to a composite microstructure, are tougher than plasma-sprayed ceramics. Plasma-sprayed ceramics are the better-performing coatings in dry particles abrasion conditions, because they never undergo abrasive grooving but only splats detachment. In particular, plasma-sprayed Al2 O3 shows the best dry particles abrasion resistance, thus, its employments are suggested under such conditions. Fracture toughness, by itself, is not a good parameter to predict plasma-sprayed ceramics dry particles abrasion resistance: the parameter KIC-L , related to the crack propagation resistance in the direction parallel to the substrate (i.e. intersplat cracking), must be introduced. A combination between KIC-L and Vickers microhardness is suggested to predict the dry particles abrasion resistance of plasma-sprayed ceramics. Cermets have similar performance to electrolytic hard chrome. In the pin-on-disk test, the ability to form a smooth and compact surface lm by local plastic deformation is the key property determining coatings performance. Plasma-sprayed Al2 O3 and Al2 O3 13%TiO2 do not form an adequately compact tribolm and therefore show unfavourable properties in terms of sample wear rate, pin wear rate and friction coefcient. Cr2 O3 , instead, has the ability to form a compact tribolm by splats plastic deformation; thus, its performances are comparable to HVOF-sprayed ceramics. Among the thick (100 m) coatings tested in this study, plasma-sprayed Cr2 O3 and HVOF-sprayed cermets are the best performing coatings under all aspects and are all candidate coatings for electrolytic hard chrome replacement. Plasma-sprayed chromia, performing similarly to cermet coatings though having potentially lower production costs, seems an interesting solution. In particular, Cr2 O3 powders, although more expensive than most ceramic ones (at least thrice more expensive than alumina), are cheaper than cermets. However, potential disadvantages, which must be considered in a proper coating choice, include the rather low deposition efciency of Cr2 O3 (45% in conventional APS [60]) and environmental problems due to possible formation of Cr6+ by in-ight thermal alteration of the coating material during spraying. Some strict rules have been recently issued in USA due to this problem [61]. Further development of this research should also involve more environmental friendly alternatives to thermally sprayed Cr2 O3 able to reproduce its excellent wear behaviour.

1314

G. Bolelli et al. / Wear 261 (2006) 12981315 [4] F. Rastegar, D.E. Richardson, Alternative to chrome: HVOF cermet coatings for high horse power diesel engines, Surf. Coat. Technol. 90 (1997) 156193. [5] M.R. Dorfman, Thermal spray materials, Adv. Mater. Process. 160 (8) (2002) 4951. [6] H. Herman, S. Sampath, R. McCune, Thermal spray: current status and future trends, in: S. Sampath, R. McCune (Eds.), Thermal Spray Processing of Materials, MRS Bulletin, 2000, pp. 1725. [7] R. Nieminen, P. Vuoristo, K. Niemi, T. M ntyl , G. Barbezat, Rolling cona a tact fatigue failure mechanisms in plasma and HVOF-sprayed WCCo coatings, Wear 212 (1997) 6677. [8] J.M. Guilemany, J.M. Miguel, S. Vizcaino, F. Climent, Role of threebody abrasion wear in the sliding wear behaviour of WCCo coatings obtained by thermal spraying, Surf. Coat. Technol. 140 (2001) 141 146. [9] L. Prchlik, S. Sampath, J. Gutleber, G. Bancke, A.W. Ruff, Friction and wear properties of WCCo and MoMo2 C based functionally graded materials, Wear 249 (2001) 11031115. [10] A. Scrivani, S. Ianelli, A. Rossi, R. Groppetti, F. Casadei, G. Rizzi, A contribution to the surface analysis and characterizatione of HVOF coatings for petrochemical application, Wear 250 (2001) 107113. [11] P.H. Shipway, L. Howell, Microscale abrasion-corrosion behaviour of WCCo hard metals and HVOF sprayed coatings, Wear 258 (2005) 303312. [12] J. Vicenzi, D.L. Villanova, M.D. Lima, A.S. Takimi, C.M. Marques, C.P. Bergmann, HVOF-coatings against high-temperature erosion (300 C) by coal y ash in thermoelectric power plant, Mater. Design 27 (2006) 236242. [13] K. Sugiyama, S. Nakahama, S. Hattori, K. Nakano, Slurry wear and cavitation erosion of thermal-sprayed cermets, Wear 258 (2005) 768775. [14] Y. Liu, T.E. Fisher, A. Dent, Comparison of HVOF and plasma-sprayed alumina/titania coatings-microstructure, mechanical properties and abrasion behaviour, Surf. Coat. Technol. 167 (2003) 6876. [15] R.S. Lima, B.R. Marple, Optimized HVOF titania coatings, J. Thermal Spray Technol. 12 (2003) 360369. [16] L. Bianchi, A. Denoirjean, F. Blein, P. Fauchais, Microstructural investigation of plasma-sprayed ceramic splats, Thin Solid Films 299 (1997) 125135. [17] R.B. Heimann, Applications of plasma-sprayed ceramic coatings, Key Eng. Mater. 122124 (1996) 399442. [18] V.V. Sobolev, J.M. Guilemany, J. Nutting, High Velocity Oxy-Fuel Spraying. Theory, Structure-Properties Relationships and Applications, Maney for the Institute of Materials, Minerals and Mining, London, 2004, p. 27. [19] G. Barbezat, Advanced thermal spray technology and coating for lightweight engine blocks for the automotive industry, Surf. Coat. Technol. 200 (2005) 19901993. [20] S. Guessasma, M. Bounazef, P. Nardin, T. Sahraoui, Wear behavior of aluminatitania coatings,analysis of process and parameters, Ceram. Int. 32 (2006) 1319. [21] J.H. Ouyang, S. Sasaki, Tribological characteristics of low-pressure plasma-sprayed Al2 O3 coating from room temperature to 800 C, Tribol. Int. 38 (2005) 4957. [22] O. Kov k, J. Siegl, J. Nohava, P. Chr ska, Youngs modulus and fatigue ar a behavior of plasma-sprayed alumina coatings, J. Therm. Spray Technol. 14 (2005) 231238. [23] Y. Xie, H.M. Hawthorne, The damage mechanisms of several plasmasprayed ceramic coatings in controlled scratching, Wear 233235 (1999) 293305. [24] Y. Xie, H.M. Hawthorne, Wear mechanism of plasma-sprayed alumina coating in sliding contacts with harder asperities, Wear 225229 (1999) 90103. [25] L.C. Erickson, H.M. Hawthorne, T. Troczynski, Correlations between microstructural parameters, micromechanical properties and wear resistance of plasma-sprayed cermaic coatings, Wear 250 (2001) 569 575. [26] R. Westergard, L.C. Erickson, N. Ax n, H.M. Hawthorne, S. Hogmark, The e erosion and abrasion characteristics of alumina coatings plasma-sprayed under different spraying conditions, Tribol. Int. 31 (1998) 271279.

Polishing generally improves the tribological performance of tested coatings; however, the friction coefcient of WCCo coating against 100Cr6 has been increased because chemical afnity troubles between steel and metal matrix have been enhanced. As-sprayed Cr2 O3 is smooth enough to cause limited wear on the counterpart and limited friction coefcients, so that, in certain instances, it could also be employed in the as-sprayed condition. This points out to the need for proper deposition parameters and the advantage in using spray powders with a quite small average particle diameter (it must not be too small, however, to prevent droplets vaporization in the plasma jet and to confer molten droplets an adequate kinetic energy to atten upon impact). The present data are also a good basis for constructing wear maps of thermally sprayed coatings, in particular for plasmasprayed Cr2 O3 , which is of high practical interest due to its favourable wear resistance. Such wear maps would be very useful in view of an increasing industrial applications in many elds where traditional plating techniques are still being employed. In particular, it is to be noticed that, in all cases considered here, the chromia coating is in the mild or ultra-mild wear regime; however, this depends on the fact that this coating has, in all cases, been able to form a compact tribolm by plastic deformation: since the tribolm is entirely ceramic (thus, intrinsically brittle), once a critical contact pressure has been exceeded, a tribolm detachment and, consequently, transition from mild to severe wear are likely to occur. The identication of this transition is of key practical importance and shall be a future development of the present work. Acknowledgements The authors thanks Centro Sviluppo Materiali S.p.A. (Roma, Italy), in particular Mr. Edoardo Severini, Mr. Francesco Barulli, Mr. Valerio Ferretti and the head of the surface engineering unit, Ing. Fabrizio Casadei, for the thermally sprayed coatings manufacturing. We are also grateful to Galvanica Nobili S.r.l. (Marano sul Panaro, MO, Italy) for hard chrome platings and to Argos S.p.A. (Monteveglio, BO, Italy) for NiP electroless platings manufacturing. A special thank to Ms. Lisa Calamai for her valuable contribution to the experimental activity. Partially supported by PRRIITT (Regione Emilia Romagna, Italy), Net-Lab Surface & Coatings for Advanced Mechanics and Nanomechanics (SUP&RMAN).

References
[1] Handbook of Thermal Spray Technology, ASM International, Materials Park, OH, USA, p. 171. [2] P.L. Ko, M.F. Robertson, Wear characteristics of electrolytic hard chrome and thermal sprayed WC10%Co4%Cr coatings sliding against AlNiBronze in air at 21 C and at 40 C, Wear 252 (2002) 880893. [3] T. Saharoui, N.-E. Fenineche, G. Montavon, C. Coddet, Alternative to chromium: characteristics and wear behaviour of HVOF coatings for gas turbine shafts repair (heavy-duty), J. Mater. Process. Technol. 152 (2004) 4355.

G. Bolelli et al. / Wear 261 (2006) 12981315 [27] J.E. Fernandez, R. Rodriguez, Y. Wang, R. Vijande, A. Rincon, Sliding wear of a plasma-sprayed Al2 O3 coating, Wear 181183 (1995) 417425. [28] K.G. Budinsky, Abrasion resistance of transport roll surfaces, Wear 181183 (1995) 938943. [29] G. Bolelli, V. Cannillo, L. Lusvarghi, S. Ricc` , Mechanical and tribological o properties of electrolytic hard chrome and HVOF-sprayed coatings, Surf. Coat. Technol. 200 (2006) 29953009. [30] S.M. Hsu, M. Shen, Wear prediction of ceramics, Wear 256 (2004) 867878. [31] K. Kato, K. Adachi, Wear of advanced ceramics, Wear 253 (2002) 10971104. [32] K. Adachi, K. Kato, N. Chen, Wear map of ceramics, Wear 203/204 (1997) 291301. [33] P. Gillespie, Electroless nickel coatings: case study, in: J.S. Burnell-Gray, P.K. Datta (Eds.), Surface Engineering CasebookSolutions to Corrosion and Wear-Related Failures, Woodhead Publishing Limited, Abington Hall, 1996, pp. 4972. [34] A.G. Evans, T.R. Wilshaw, Quasi-static solid particle damage in brittle solids. I. Observations analysis and implications, Acta Metall. 24 (1976) 939956. [35] C.B. Ponton, R.D. Rawlings, Vickers indentation fracture toughness test. Part 1: Review of literature and formulation of standardised indentation toughness equations, Mater. Sci. Technol. 5 (1989) 865872. [36] C.B. Ponton, R.D. Rawlings, Vickers indentation fracture toughness test. Part 2: Application and critical evaluation of standardised indentation toughness equations, Mater. Sci. Technol. 5 (1989) 961976. [37] Y. Liu, T.E. Fischer, A. Dent, Comparison of HVOF and plasma-sprayed alumina/titania coatingsmicrostructure, mechanical properties and abrasion behaviour, Surf. Coat. Technol. 167 (2003) 6876. [38] E. Lopez-Cantera, B.G. Mellor, Fracture toughness and crack morphologies in eroded WCCoCr thermally sprayed coatings, Mater. Lett. 37 (1998) 201210. [39] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments, J. Mater. Res. 7 (1992) 15641583. [40] http://www.matweb.com, oxide ceramics database. [41] J. Alcal` , F. Gaudette, S. Suresh, S. Sampath, Instrumented spherical microa indentation of plasma sprayed coatings, Mater. Sci. Eng. A 316 (2001) 110. [42] A. Kulkarni, J. Gutleber, S. Sampath, A. Goland, W.B. Lindquist, H. Herman, A.J. Allen, B. Dowd, Studies of the microstructure and properties of dense ceramic coatings produced by high-velocity oxygen-fuel combustion spraying, Mater. Sci. Eng. A 369 (2004) 124137. [43] A. Kulkarni, Z. Wang, T. Nakamura, S. Sampath, A. Goland, H. Herman, J. Allen, J. Ilavsky, G. Long, J. Frahm, R.W. Steinbrech, Comprehensive microstructural characterization and predictive property modeling of plasma-sprayed zirconia coatings, Acta Mater. 51 (2003) 2457 2475. [44] J. Zhang, V. Desai, Evaluation of thickness, porosity and pore shape of plasma sprayed TBC by electrochemical impedance spectroscopy, Surf. Coat. Technol. 190 (2005) 98109.

1315

[45] A.C. Fox, T.W. Clyne, Oxygen transport by gas permeation through the zirconia layer in plasma sprayed thermal barrier coatings, Surf. Coat. Technol. 184 (2004) 311321. [46] A.J. Slifka, B.J. Filla, J.M. Phelps, G. Bancke, C.C. Berndt, Thermal conductivity of a zirconia thermal barrier coating, J. Therm. Spray Technol. 7 (1998) 4346. [47] A. Kulkarni, A. Vaidya, A. Goland, S. Sampath, H. Herman, Processing effects on porosity-property correlations in plasma-sprayed yttriastabilized zirconia coatings, Mater. Sci. Eng. A 359 (2003) 100 111. [48] K. Ramachandran, V. Selvarajan, P.V. Ananthapadmanabhan, K.P. Sreekumar, Microstructure, adhesion, microhardness, abrasive wear resistance and electrical resistivity of the plasma-sprayed alumina and aluminatitania coatings, Thin Solid Films 315 (1998) 144152. [49] G. Bolelli, R. Giovanardi, L. Lusvarghi, T. Manfredini, E. Soragni, A. Zanichelli, Caratterizzazione microstrutturale, tribologica ed elettrochimica di varie tipologie di rivestimenti in cromo duro, in: XXX Congresso Nazionale AIM, Vicenza, Italy, 2004. [50] G.B. Stachowiak, G.W. Stachowiak, The effects of particle characteristics on three-body abrasive wear, Wear 249 (2001) 201207. [51] J.D. Gates, Two-body and three-body abrasion: a critical discussion, Wear 214 (1998) 139146. [52] B. Bhushan, Principles and Applications of Tribology, John Wiley and Sons, New York, 1999, p. 509. [53] M. Factor, I. Roman, Microhardness as a simple means of estimating relative wear resistance of carbide thermal spray coatings: Part 1. Characterization of cemented carbide coatings, J. Therm. Spray Technol. 11 (2002) 468481. [54] M. Factor, I. Roman, Microhardness as a simple means of estimating relative wear resistance of carbide thermal spray coatings: Part 2. Wear resistance of cemented carbide coatings, J. Therm. Spray Technol. 11 (2002) 482495. [55] H.-S. Ahn, O.-K. Kwon, Tribological behaviour of plasma-sprayed chromium oxide coating, Wear 225229 (1999) 814824. [56] Q. Yang, T. Senda, A. Ohmori, Effect of carbide grain size on microstructure and sliding wear behaviour of HVOF-sprayed WC12%Co coatings, Wear 254 (2003) 2334. [57] R. Schwetzke, H. Kreye, Microstructure and properties of tungsten carbide coatings sprayed with various high-velocity oxygen fuel spray systems, J. Therm. Spray Technol. 8 (3) (1999) 433439. [58] P.H. Shipway, D.G. McCartney, T. Sundapresert, Sliding wear behaviour of conventional and nanostructured HVOF-sprayed WCCo coatings, Wear 259 (2005) 820827. [59] M. Scherge, D. Shakhvorostov, K. P hlmann, Fundamental wear mechao nism of metals, Wear 225 (2003) 395400. [60] Sulzer-Metco web site, http://www.sulzermetco.com. [61] State of California Air Resources Board, Resolution 04-44, December 9, 2004; available on-line at: http://www.arb.ca.gov/regact/thermspr/ res0444.pdf.

S-ar putea să vă placă și