Sie sind auf Seite 1von 9

Journal

J. Am. Ceram. Soc., 94 [12] 44264434 (2011) DOI: 10.1111/j.1551-2916.2011.04899.x 2011 The American Ceramic Society

High-Temperature Tribological and Self-Lubricating Behavior of Copper Oxide-Doped Y-TZP Composite Sliding Against Alumina
Mahdiar Vale,, Matthijn de Rooij, Dirk J. Schipper, and Louis Winnubst

Surface Technology and Tribology Group, Faculty of Engineering Technology, University of Twente, 7500 AE Enschede, The Netherlands

Inorganic Membrane Group, Faculty of Science and Technology, University of Twente, 7500 AE Enschede, The Netherlands eects of dierent yttria contents and suggested that higher fracture toughness tended to increase the wear resistance for yttria-doped zirconia ceramics. Lee et al.8 presented a series of three-dimensional wear maps for Y-TZP under dry and lubricated conditions. They observed a sudden increase in wear as a function of load and velocity. The mild wear mechanism is ascribed to plastic deformation and microfracture. Severe wear is dominated by brittle fracture and third body abrasion. Rainforth9 considered the role of transformation toughening in Y-TZP and correlates the poor wear properties of zirconia to the low thermal conductivity of this material which results in a substantial ash temperature rises at the surface, thus leading to softening of the material. It has been reported that Y-TZP shows high wear and friction at high temperature in comparison with room temperature.11 Wear studies of zirconia matrix self-lubricating composites have been reported in the literature.1214 In addition, a mechanical model has been recently introduced by Song et al.15 to predict the thickness of a self-lubricating layer in ceramic composite materials containing a solid lubricant. Solid lubricants, such as graphite, molybdenum disulde, and some soft metals, cannot resist temperatures above 500C due to thermal decomposition or oxidation. Only a few solid lubricants are suitable for high temperature conditions, which include some stable uorides and lubricious oxides.16,17 Among them, copper oxide as a solid lubricant shows a promising reduction in friction, not only as a thin coating18 but also when it is embedded in a zirconia matrix.19 Recently, it has been observed that CuO/Y-TZP composites show low friction when sliding against alumina at room temperature.12,19 Details of the mechanism are discussed elsewhere.20 Prakash and Celis21 have investigated lubricity of a wide variety of solid oxides at high temperatures, based on a polarizability approach. According to their approach CuO shows low friction (l = 0.180.22) at T = 584C. Therefore, it is interesting to investigate the wear of ceramic composite doped with copper oxide at the mentioned temperature. It is noteworthy to mention that none of these works investigated the friction and wear mechanism of CuO-zirconia at high temperatures. In this study, high temperature friction and wear of CuO TZP system against an alumina ball are systemically investigated. The results are compared with TZP without CuO. The wear mechanism was investigated with interference microscopy, SEM, EDS, and XPS.

The tribological behavior of 5 wt% copper oxide-doped tetragonal zirconia polycrystal composite has been investigated while it slides against an alumina counterface under high temperature conditions. The eects of load (1, 2.5, and 5 N) and velocity (0.05 and 0.1 m/s) on the wear mechanism have been investigated. The results were compared with undoped zirconia at 595C. A coecient of friction (COF) of 0.35 and a specic wear rate less than 106 mm3/Nm were obtained at 595C when copper oxide was added to zirconia. Further, it has been observed that a self-lubricating layer is formed at the interface. Scanning electron microscopy and X-ray photoelectron spectroscopy have been used to investigate the formation of the self-lubricating copper-rich layer at the interface between the disk and counterface materials. The formation of a self-lubricating layer, as well as the wear mechanisms at dierent operational conditions (load and velocity) are discussed. It appears that plastic deformation of copper-rich phase at higher temperature is responsible for the decrease in friction and wear.

I.

Introduction

ERAMICS based on yttria-stabilized tetragonal zirconia polycrystals (Y-TZP) are suitable materials for tribological applications due to their excellent wear and mechanical properties. For instance, its wear-resistant applications include: automotive engines, conveyor belt scraper blades for the mining industry, and dies for hot extrusion of copper.1 However, investigation of the friction and wear of this material at elevated temperature has revealed that it exhibits a high COF under dry conditions.2,3 To apply ceramic materials in tribological applications, the COF must be reduced. As a liquid lubricant is not feasible for high temperature applications, a solid lubricant is a logical choice. In general, solid lubricants which form a thin coating by brushing, sputter coating, or plasma spraying can reduce friction. However, they are only ecient as long as the initial lms are present on the surface. To have continuous solid lubricant lms at the interface, an eective approach is to incorporate solid lubricants into the ceramic matrix as a second phase. The second phase generates a soft layer at the surface which can restore itself during sliding.4 Numerous studies have been done on the wear of Y-TZP at room temperature and at high temperatures.510 Fischer et al.5 investigated the wear behavior of yttria-doped zirconia materials under various environments. They studied the

II.
R. Cutlercontributing editor

Experimental Procedure

Manuscript No. 29669. Received April 30, 2011; approved September 14, 2011. Author to whom correspondence should be addressed. e-mail: m.vale@utwente. nl, mahdiyar59@yahoo.com

(1) Material Preparation Disk of 8 mol% (5 wt%) CuO-doped Y-TZP (5CuOTZP) were prepared using 3Y-TZP (3 mol% yttria-doped tetragonal zirconia polycrystals: TZ3Y, Tosoh, Tokyo, Japan) and CuO (Aldrich, Steinheim, Germany) powders as the starting 4426

December 2011

High-Temperature Tribological Behaviour of CuO Doped Y-TZP Composite

4427

materials. The particle size of 3Y-TZP and CuO powder were 64 nm and 74 lm, respectively. The processing route and mechanical properties were described in detail elsewhere.20,22 Cold-pressed 5CuOTZP samples were sintered at 1500C for 8 h with heating and cooling rate of 2C/min. The sintered density was measured by immersion in mercury using the Archimedes method. The sintered disks were polished to a surface roughness (Ra) of <50 nm using a diamond paste. The polished disks were ultrasonically cleaned in ethanol and then annealed at 850C for 2 h. For comparison, undoped 3Y-TZP and 0.8 mol% (0.5 wt%) 0.5CuO TZP were prepared and sintered at 1400C and 1500C for 2 h, respectively.

(2) High-Temperature Tribological Experiments High-temperature tribological experiments were conducted on a pin on disk tribometer (CSEM, Neuchatel, Switzerland) capable of testing up to 800C. Commercial alumina balls with mirror-polished surfaces (GIMEX Technische Keramiek B.V., Geldermalsen, the Netherlands) and a diameter of 10 mm were used as pin. Before the sliding tests, both ball and disk were ultrasonically cleaned with isopropanol for 30 min and rinsed with deionized water and then dried in oven at 120C for 24 h. The applied tests conditions were 1, 2.5, and 5 N load at velocities of 0.05 and 0.1 m/s; the radius of the wear track of 1012 mm; and a total sliding distance of 1 km at 595C. To minimize frictional heating, the experiments have been conducted at low load and velocity. All experiments were repeated at least three times to check the reproducibility of the results. Surface roughness was examined using interference microscopy. The wear volume of the disks was determined by measuring their mass before and after testing using a mass balance with a resolution of 0.01 mg. The mass dierence was then normalized by dividing the volume loss by the applied load and the sliding distance to obtain specic wear rate. As the wear volume of 5CuOTZP was too low in the case of 1 N test condition, it was determined using a surface prolometry scan. (3) Material Characterization The microstructure and wear mechanism were investigated using a scanning electron microscope (JSM6400 and JCM5000; JEOL, Tokyo, Japan) equipped with energy dispersive spectroscope (EDS). The phase identication of zirconia composites was performed by X-ray diraction (Philips, Xpert APD, PANalytical, Almelo, The Netherlands). X-ray photoelectron spectrometry (XPS) was used to study the chemical changes in the wear track, unworn track, and wear scar of the ball (Model Quantera SXM; Physical Electronics, Chanhassen, MN) using Al Ka(1486.6 eV) excitation source with a monochromator). The hardness was measured by means of a Vickers indentation test at a load of 10 N on polished samples. Five indentations were made for each material. The diagonal lengths of the indentation impressions were determined using an optical microscope with Normaski interference coupled with a video micrometer. The fracture toughness was obtained using an indentation-strength method.23 It is worth noting that this method gives only relative information about fracture toughness of the two materials investigated in this work, as indentation fracture toughness is not an absolute value. III. Results

well as monoclinic zirconia phases. Based on XRD patterns, the volume fraction of monoclinic and tetragonal phases are calculated to be 70% and 30%, respectively, using the method of Toraya et al.24 Presence of CuO and Y2Cu2O5 phases could not be identied by means of XRD due to their low content and due to interference with diraction pattern of tetragonal and monoclinic zirconia. Figure 1 shows a typical microstructure of polished and thermally etched surface of 5CuOTZP composite. The SEM micrograph [Fig. 1(a)] reveals that 5CuOTZP has an average grain size around 2 lm. The undoped 3Y-TZP shows an average grain size of 0.5 lm. The larger grain size in CuO TZP is ascribed to a higher sintering temperature and the longer holding time during sintering, as well as to the presence of CuO.22 Figure 1(b) is the EDS spectrum of 5CuO TZP indicating Cu-rich grains and grains which do not contain Cu above the detection limit of EDS. Based on the SEM observation, as well as on the EDS analysis, one may conclude that copper-rich grains can be formed after sintering when 5 wt% CuO is added to 3Y-TZP. Ran et al. found similar results using TEM/EDS.22 Table I shows the mechanical properties of the materials used in this study. The 5CuOTZP shows slightly lower fracture toughness as compared to TZP. This can be attributed to the presence of a copper-rich grain-boundary phase which will be discussed later.

(2) Friction Behavior Table II shows the results of COF as function of test conditions for 5CuOTZP and TZP. The average COF (l) varies between 0.35 and 0.56 for 5CuOTZP. In contrast, undoped TZP shows a higher average COF under the same conditions, varying from 0.7 to 0.81, which is comparable with that reported in literature.3,6 Interestingly, the addition of copper oxide to TZP results in a 35%50% reduction in COF. Figure 2 exhibits the friction traces of disks with dierent levels of CuO when sliding against alumina ball at 595C. According to Fig. 2, both 3Y-TZP and 0.5CuOTZP show a high COF with an average value of 0.7 and 0.71, respectively. In contrast, 5CuOTZP shows signicantly lower COF with an average value of 0.35 0.05. The lower COF

(a)

(b)

(1) Microstructure and Mechanical Properties The relative densities of sintered 3Y-TZP and 5CuOTZP composites were 98% and 97%, respectively. The XRD analysis of 5CuOTZP revealed the presence of tetragonal, as

Fig. 1. Micrograph of polished and thermally etched 5CuOTZP using (a) SEM and (b) corresponding EDS analysis (note that the arrows indicate location at which EDS analysis was performed).

4428

Journal of the American Ceramic SocietyVale et al.


Table I. Mechanical Properties of the Materials Used in This Study
Bending strength (MPa) Fracture toughness (MPam1/2) Grain size (lm)

Vol. 94, No. 12

Material

Hardness (GPa)

Elastic modulus (GPa)

CLA surface roughness (nm)

Alumina 3Y-TZP 5CuOTZP

20 0.8 13 0.4 9.5 0.5

320 210 205

340 450 20 310 10

3.4 0.4 4.2 0.2 3.5 0.2

5 0.45 2

15 25 25

Bending strength data taken from Ran et al.22

Table II.
Material

Maximum Hertzian Contact Pressure (pmax), Flash Temperature (Tf), Tensile Stresses(rmax), as well as Wear Data in This Study for Doped and Undoped 3Y-TZP Tested at 595C
v (m/s) F (N) l () k(9106) (mm3 N1 m1) pmax (GPa) rmax (GPa) Tf (C)

5CuOTZP

0.05 0.1

3Y-TZP

0.05 0.1

1 2.5 5 1 2.5 5 1 2.5 5 1 2.5 5

0.35 0.53 0.56 0.4 0.55 0.53 0.7 0.78 0.79 0.76 0.81 0.77

0.05 0.01 0.01 0.05 0.01 0.01 0.03 0.01 0.01 0.01 0.01 0.01

0.4 360 460 0.5 895 1425 586 307 250 599 359 436

0.005 24 40 0.004 50 30 80 7.25 18 2.5 8.5 62

0.52 0.71 0.89 0.52 0.71 0.89 0.52 0.71 0.9 0.52 0.71 0.9

0.37 0.73 0.96 0.42 0.76 0.92 0.68 1.03 1.32 0.74 1.06 1.3

695 743 751 807 897 886 801 827 830 1043 1072 1048

1.2 1
Friction coefficient

a)3Y-TZP b) 0.5 CuO-TZP c) 5 CuO-TZP

1 0.9 0.8 Friction coefficient 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0 100 200 300 400 500 Temperature(oC) 600 700

CuO-TZP TZP

0.8 0.6 0.4 0.2 0 0 200 400 600 800

a) b) c)

1000

800

Sliding distance(m)

Fig. 2. Coecient of friction (COF) as a function of sliding distance for monolithic and TZP composites sliding against alumina ball (F = 1 N, v = 0.05 m/s, T = 595C).

Fig. 3. Friction coecient as a function of test temperature for 5CuOTZP and TZP against alumina ball (F = 1 N, v = 0.05 m/s).

of 5CuOTZP composite compared with TZP might be attributed to the lubrication eect of the second phase. In addition, a lower COF for 5CuOTZP as compared to 0.5CuOTZP suggests that the amount of second phase has a key role in friction and wear behavior of CuOTZP. To analyze the reason for decrease in the friction coecient when CuO is added to Y-TZP, friction measurements were done at dierent temperatures. Figure 3 shows the friction coecient as a function of test temperature for undoped TZP and 5CuOTZP. Both materials show the same trend up to 200C while 5CuOTZP starts to drop in COF level drastically above 200C. It is clear from this plot that the 5CuOTZP composite shows signicantly reduced values of friction compared to TZP at higher temperatures. From this result, three friction regions can be distinguished for 5CuO TZP: at room temperature, intermediate temperature (25C < T < 595C), and high temperature (T > 595C) regimes. At room temperature and above 595C, 5CuOTZP shows a low friction coecient, whereas at intermediate temperatures, a high friction coecient is observed. Jahanmir et al.25,26 reported a similar observation for high-temperature wear of alumina and silicon nitride.

(3) Wear Behavior Specic wear rate (k) data of both TZP and 5CuOTZP are included in Table II. The 5CuOTZP disk shows extremely low wear rate (0.43 9 106 mm3/Nm) at applied load of 1 N and at both tested velocities of 0.05 and 0.1 m/s. This is independent of the velocity. At higher loads, the specic wear rate signicantly rises for 5CuOTZP disk material when compared with TZP. Based on the COF and specic wear rate data, mild (specic wear rate < 106) and severe wear (specic wear rate > 106) have been observed for 5CuOTZP at dierent loads. Mild wear occurred at low load (F 1 N). At higher loads (F > 1 N) where surface microcracking and chipping manifest, severe wear is responsible for material loss. (4) Microstructural Studies in Mild Wear The SEM and EDS analyses were used to elucidate the wear mechanism in 5CuOTZP and to compare the wear mechanism with undoped TZP. Figure 4 shows the SEM micrographs of unworn and worn 5CuOTZP at 595C. The polished sample [Fig. 4(a)] shows (polishing) lines and porosity. However, the wear track is quite smooth and the

December 2011
(a)

High-Temperature Tribological Behaviour of CuO Doped Y-TZP Composite


(b)

4429

(c)

(d)

(e)

Fig. 4. SEM micrographs of the (a) unworn and (b) smooth worn 5CuOTZP after pin on disk test (F = 1 N, v = 0.05 m/s, T = 595C), higher magnication micrograph (c) of a polished 5CuOTZP surface and (d, e) after wearing test (arrows show wear debris and patchy layer).

polishing lines almost disappear after the sliding experiment [Fig. 4(b)]. To nd out more details, higher magnication SEM images have been shown in Figs. 4(c)(e). According to Figs. 4(d) and (e), signicant amounts of tiny wear debris with a diameter of <100 nm are entrapped in holes after sliding test. This indicates that the original pores in 5CuOTZP [Fig. 4(e)] can serve as sinks for the submicrometer wear debris generated through deformation and microfracture. The size of the wear debris particles and the smoothness of the surface after the wear test imply that wear occurs on a scale which is signicantly smaller than the grain size. Furthermore, near the pores, a patchy layer can be distinguished [Figs. 4(d) and (e)]. In addition, surface roughness measurements of disks using AFM indicate a smooth wear track with similar roughness to that of the polished sample. Figure 5 shows a SEM micrograph of a worn 3Y-TZP surface. It is obvious that the wear track on an undoped sample, tested under the same conditions show a rough appearance which is characterized by severe plastic deformation and microcrack generation in the wear track. To obtain more information about the mechanism for the low friction of the alumina/5CuOTZP systems, SEM micrographs of alumina balls were also studied. Figure 6 depicts a wear scar of an alumina ball tested against 5CuOTZP and

Fig. 5. SEM micrograph of worn surface of 3Y-TZP tested at F = 1 N, v = 0.05 m/s, T = 595C (arrow shows the sliding direction).

3Y-TZP. The micrograph 6a clearly indicates a smooth wear scar and accumulation of wear debris ahead and beside the wear scar. Micrograph 6b illustrates a rough wear scar and

4430

Journal of the American Ceramic SocietyVale et al.

Vol. 94, No. 12

extended worn scar of an alumina ball sliding against 3Y-TZP, indicating a more severe wear. It is noteworthy to mention that the diameter of wear scar on the alumina ball
(a)

sliding against 3Y-TZP (880 lm) is ~3.7 times larger when compared to alumina ball sliding against 5CuOTZP (235 lm). This, results in a wear rate increase by two orders of magnitude for the counterface and suggests that 5CuO TZP causes less wear of the alumina counterface. An EDS composition mapping analysis combined with SEM was conducted to detect the elemental distribution of the accumulated wear debris on the alumina ball (Fig. 7). As can be seen from the image, a copper-rich layer was formed at the edge of the ball.

(b)

Fig. 6. SEM micrograph of alumina ball sliding against (a) 5CuO TZP (b) TZP, both tested at F = 1 N, v = 0.05 m/s, T = 595C (arrow shows the sliding direction and dotted circle shows the contact area).

(5) XPS Study of a Worn Surface at 595C Figure 8 shows an XPS spectrum of the copper-rich layer formed on the alumina ball. The XPS spectrum shows two main peaks: (i) the core level Cu 2p3/2 XPS at 934.5 eV accompanied by satellite peaks toward a higher binding energy of ~940.8 and 943 eV, (ii) the core level Cu 2p3/2 XPS transition at 932.2 eV. By comparing the energy values in the Cu(2p) spectra with the data in the Handbook of X-ray Photoelectron Spectroscopy and literature,27,28 it may be concluded that pure CuO (with peaks at BE = 934.5 eV) and Cu2O (with peaks at BE = 932.2 eV) are present in the lm. This indicates that the copper-rich layer is a mixture of CuO and Cu2O. The satellite feature observed in XPS spectrum is the characteristic of materials, such as copper halides or CuO, having d 9 conguration in the ground state.27 However, the d shell of Cu2O is full d10, so that screening via a charge transfer into the d state is not possible. This explains the absence of a satellite peak in Cu2O. To further analyze, to what extent copper oxide phase is responsible for friction reduction, depth prole using XPS was conducted on the 5CuOTZP wear surface. Figure 9 shows depth proles for both inside and outside the wear track as a function of sputtering time. The results show copper enrichment inside the wear track which decreases to a steady state value after 9 min of sputtering. Based on the sputtering time, the thickness of the layer can be estimated to be around 60 nm. Furthermore, the same analysis was performed on the samples tested at different temperatures. The results are shown in Table III.

Fig. 7. EDS composition mapping of alumina ball sliding against CuOTZP disk at 595C (a) Backscatter SEM photograph of EDS-mapped area, (b) O distribution, (c) Al distribution, and (d) Cu distribution.

December 2011

High-Temperature Tribological Behaviour of CuO Doped Y-TZP Composite


(a)

4431

(b)
Fig. 8. Cu 2p XPS spectra of layer formed on alumina ball sliding against 5CuOTZP at 595C, F = 1 N, v = 0.05 m/s.

Inside wear track Outside wear track

5 Cu-atomic concentration (%)

Fig. 10. SEM micrographs indicating severe wear in (a) 5CuOTZP (b) 3Y-TZP wear track due to testing at T = 595C, F = 5 N, v = 0.1 m/s (arrows show the sliding direction).

0 0 2 4 6 8 10 Sputter time (min)

Fig. 9. Depth composition prole of 5CuOTZP disk sliding against alumina ball (F = 1 N, v = 0.05 m/s, T = 595C).

Table III.

Element Ratio of Cu/Zr and Al/Zr as a Function of Depth Prole Using XPS
200C 400C Inside track Outside track Inside track 595C Outside track Inside track

Ratio

Outside track

Cu/ Zr Al/ Zr

0.02 0

0.09 0.12

0.07 0

0.08 0.11

0.03 0

0.12 0.13
Fig. 11. 5CuOTZP cross-section showing a lateral crack beneath the wear track (tested at 595C, F = 5 N, v = 0.1 m/s). The arrow points to the surface edge and sliding direction is perpendicular to the image plane.

According to Table III, it is obvious that the Cu/Zr ratio increases after a sliding test in the wear track regardless of temperature.

(6) Severe Wear and Abrasive Wear Figure 10 shows severe wear at higher load and velocity. It is re-iterated herein that 5CuOTZP shows much larger wear as compared to 3Y-TZP at higher loads. The wear track of 5CuOTZP [Fig. 10(a)] suggests that material removal occurs by abrasion and microfracture at higher loads. The worn surface of 3Y-TZP [Fig. 10(b)] is dierent from CuO-doped TZP and severe plastic deformation, grain fragmentation, delamination, as well as formation of microcracks perpendic-

ular to the sliding direction are visible, while no grooves are observed. To provide more proof on abrasive wear, a 5CuOTZP disk with severe wear track was cut transverse to the track using a high-speed cutter. The cross-section was studied using SEM. Figure 11 shows the formation of a lateral crack beneath the wear track. It is most likely that these lateral cracks intersect, spall, and cause signicant wear of 5CuOTZP at high loads, as observed in other studies.29 This observation suggests that generated wear debris can act as abrasive particles and increase the wear rate of 5CuO TZP.

4432 IV.

Journal of the American Ceramic SocietyVale et al. Discussion

Vol. 94, No. 12

(1) Microstructural Aspects of 5CuOTZP The results of this study have shown that tribological properties of 3Y-TZP sliding against alumina at elevated temperature can be improved by addition of CuO, enabling it a self-lubricating ceramic composite. It is conrmed from the analysis of friction traces that CuO has a profound inuence on the average COF of 5CuOTZP. Therefore, it is of great importance to know how CuO interacts with 3Y-TZP (Fig. 2). It is well-known that addition of CuO to 3Y-TZP results in the formation of Y2Cu2O5 with a Tm = 1110C.3032 During sintering of CuO and 3Y-TZP, several reactions, including the formation of Y2Cu2O5 at and above 850C, can occur.3032 It is important to recall herein that XRD analysis of 5CuOTZP reveals that 70% monoclinic and 30% tetragonal zirconia phases are present after sintering. Ran et al.33 and Lemaire et al. 31 found that even a small amount (0.5 wt%) of CuO in 3Y-TZP results in transformation of tetragonal phase into ~ 55% monoclinic phase. This transformation has been explained as follows33: (1) migration of yttria toward grain boundary at high temperature, (2) reaction between yttria and CuO, resulting in the formation of Y2Cu2O5, (3) enhanced kinetic reaction by aforementioned reaction resulting in a decrease in kinetic barrier energy for tetragonal to monoclinic phase transformation and therefore destabilization of the tetragonal phase during cooling. Unfortunately, it was not possible to identify CuO and Y2Cu2O5 phases in 5CuOTZP by XRD due to the small volume fraction of second phases and overlap with diraction lines from tetragonal and monoclinic zirconia phases. However, the presence of a copper-rich crystalline phase has been conrmed by TEM/EDS analyses in the previous study.22 (2) Thermal and Mechanical Aspects of Wear The friction and wear results in this study reveal that 3Y-TZP shows a dierent wear mechanism compared to 5CuOTZP at 595C when sliding tests performed against an alumina counterface. In case of 5CuOTZP, a distinct dierence in wear mode is observed as a function of load, whereas for single phase 3Y-TZP, the wear mode is independent of load. Various studies on tribological behavior of ceramics have shown that wear mechanisms depend on contact conditions.25,34 Therefore, it is important to assess contact stresses and ash temperatures and compare 3Y-TZP and 5CuOdoped TZP. To estimate the ash temperature, the following analytical model has been used35:
Tf Tb0   lFv 1 Ar K1 =l1f K2 =l2f Ar Tb T0 An (1)

the hardness of the softer material. The tribocouples in this study have a nominal contact area of 2.83 9 109 m2 at a load of 1 N. Using KTZP = 2.2 W(mK)1 and KAl2 O3 13 W(mK)1 at 600C,36,37 by assuming that KCuO = 17.2 W(mK)1 at 600C is the same at room temperature as also measured by Schreck and Rohde38 for CuOLTCC (low-temperature co-red ceramic), one can calculate the K5CuOTZP = 3.4 W(mK)1 using the rule of mixtures. The results of the estimated ash temperatures are shown in Table II. Addition of CuO to 3Y-TZP increases the thermal conductivity and decreases the COF which results in signicantly lower ash temperatures in the mild wear regime (F = 1 N) as compared to 3Y-TZP. The CuO would promote faster dissipation of heat, generated by the sliding contact at the interface. These result in moderate thermal severity, and diminish the contribution of thermal stresses to the wear process as compared to 3Y-TZP. With respect to the mechanical aspects, the initial maximum Hertzian contact pressure pmax and maximum tensile stresses rmax induced by friction at the back of the contact are calculated using the following equations39: 3F 2pa2

pmax

(4)

rmax

  1 2m 4 m pmax pl 3 8

(5)

Tb0 Tb

(2)

Tb T0

  lFv 1 An K1 =l1b K2 =l2b

(3)

where F is the load (N), a is the initial Hertzian contact radius (m), m is Poissions ratio, and l is COF. The estimated contact pressures are also summarized in Table II. It is clear from Table II that the calculated initial Hertzian pressures at load 1 N (mild wear regime) are almost similar in magnitude (~520 MPa) for both materials. The initial Hertzian contact diameter is also calculated to be 60 lm. Based on SEM micrographs of wear scars (width of 235 and 880 lm for 5CuOTZP and 3Y-TZP, respectively, tested at F = 1 N, v = 0.05, and T = 595C), nal contact stresses are calculated to be 23 and 1.64 MPa, respectively. The higher nal contact stresses for 5CuOTZP indicate that 5CuO TZP is able to carry higher contact pressures and withstand (~1.8 times) higher tensile stresses as compared to 3YTZP. The SEM micrographs [Figs. 5 and 10(b)] of 3Y-TZP have shown that cracks perpendicular to the sliding direction are formed. Therefore, it is expected that the combination of thermal and mechanical stresses exceeds Kc of the material and microfracture occurs and generate intergranular cracks. One can conclude that the addition of CuO to 3Y-TZP not only results in a lower friction because of the presence of the soft layer, but also enhances the wear resistance of TZP. Finally, it can be re-iterated that 5CuOTZP shows a mild to severe wear transition with increasing load. The initial Hertzian contact pressure at 5 N is similar for the two materials in this study. Based on measured specic wear rate at corresponding test condition (Table II), 5CuOTZP shows a higher specic wear rate compared to TZP. As contact pressure increases, formation and intersection of subsurface cracks (Fig. 11) result in signicant material loss. This may be attributed to the slightly lower hardness and fracture toughness, as well as, to the presence of a grain-boundary phases.

where subscript 1 refers to the ball and subscript 2 to the disk. The An and Ar are the nominal and real contact areas (m2); l1b and l2b, as well as lb, and lf are equivalent heat diusion distances (m); F, normal load (N); v, sliding velocity (m/ s); l, the COF (), and K, the thermal conductivity (W(mK)1). Furthermore, T0, Tb, and Tf are the heat sink temperature, bulk and ash temperatures (K), respectively. The real contact area is estimated by Ar = F/H where H is

(3) Wear Mechanism Based on SEM observation (Fig. 4), it is clear that a smooth layer is formed by reattachment of submicrometer wear debris on the wear track when an alumina ball is sliding against 5CuOTZP at 595C. The results support other studies25,40 which show that the formation of a smooth layer at high temperature results in lower friction. Depth prole analysis using XPS on the wear track indicates that a thin layer,

December 2011
(a)
P

High-Temperature Tribological Behaviour of CuO Doped Y-TZP Composite

4433

(b)

Squeezed out copper rich phases

Wear debris

According to Fig. 3, both 3Y-TZP and 5CuOTZP show low friction at room temperature. A low COF for 5CuOTZP and 3Y-TZP at room temperature and at low loads has been recently discussed and attributed to the formation of a soft aluminum hydroxide layer.20 When the temperature increases to 200C, both materials clearly show a high COF. This may be attributed to the instability of aluminum hydroxide above room temperature.25 However, 5CuOTZP shows a drop in friction above 200C. It is worth recalling that second (Cu)phase oxides become soft in the temperature range of 450C 920C. Hence, a lubrication eect becomes dominant for 5CuOTZP when the temperature is increased above 200C. It emerges from results of this study that addition of CuO to 3Y-TZP enhances the wear resistance of the composite at high temperature. One of the main strategies in tribomaterials is that it should be designed with the primary object of lowering friction. The production of self-lubricating composites with a soft second phase thus makes it possible to design materials with enhanced tribological properties as compared to monolithic materials.

(c)

V.

Conclusion

Based on wear and friction studies on CuOTZP composites and 3Y-TZP at high temperatures, the following conclusions can be made: 1. The addition of 5 wt% CuO to 3Y-TZP results in the formation of copper-rich phases in the material which becomes soft at elevated temperature. 2. A self-lubricating soft layer is formed at temperature above 200C at the interface while alumina is sliding against 5CuOTZP. 3. An average COF of 0.35 0.05 and a specic wear rate less than 106 mm3/Nm was measured for 5CuO TZP tested at 595C. The maximum Hertzian pressure was 550 MPa. The SEM and XPS analyses prove that the layer is a copper-rich phase. Measurements conducted under the same condition for 3Y-TZP show high friction (COF = 0.69) and wear. 4. At a temperature of 595C, the specic wear rate of alumina ball sliding against 5CuOTZP is decreased up to two orders of magnitude as compared to 3YTZP. The formation of a soft lubricating lm enriched with copper at the interface can improve the wear resistance and change the wear mechanism from brittle fracture to plastic deformation.

(d)

Fig. 12. Schematic representation of the wear mechanism in dierent sliding cycles (a) start contacting, (b) formation of wear debris, (c) plastic deformation of debris and deformation of patchy layer, and (d) formation of transfer layer on ball.

mainly consisting of copper oxide, with a thickness of ~60 nm is present on the wear track. In addition, XPS and EDS observation, indicating the formation of the Cu2+ and Cu1+ copper-rich layer on the counterface (Figs. 7 and 8), suggest that the copper-rich phases can act as a solid lubricant at the interface. Based on the above observation and discussion about second phase, one can conclude that a mixture of two phases (CuO and Y2Cu2O5) is most likely present when 5% CuO is added to 3Y-TZP. In consideration of tribological behavior, both copper oxide and Y2Cu2O5 are soft at high temperature conditions. It is well-known that oxides become soft above their ductile-to-brittle transition temperature41 which is typically ~0.40.7 Tm, where Tm is melting point.42 It is expected that Y2Cu2O5 and CuO, with Tm values of about 1110C and 1300C, start to soften between 450C800C and 520C920C, respectively. Furthermore, it is known that Cu2O shows extensive plastic deformation at 600C.43 Hence, one can conclude that the copper oxide-rich phases can be squeezed out as a result of plastic deformation, and form a smooth layer. This layer decreases the COF in further sliding, shears and accumulates at the edge of wear scar of the alumina ball [Fig. 7(a)]. The process of layer formation is repeated and in this way, the layer is restored at the interface. A schematic representation of this mechanism is shown in Fig. 12.

Acknowledgments
The authors thank S. Jahanmir for his fruitful discussions. Thanks are also to B. Pathiraj for his comments on the manuscript, and L. Vargas and E.G. de Vries for assistance with SEM and hardness measurements, respectively. G. Kip is appreciated for his XPS analysis. Funding of this work was provided by the Dutch governmental program IOP Self Healing Materials.

References
R. H. J. Hannink, M. J. Murray, and H. G. Scott, Friction and Wear of Partially Stabilized Zirconia: Basic Science and Practical Applications, Wear, 100, 35566 (1984). 2 G. W. Stachowiak and G. B. Stachowiak, Unlubricated Friction and Wear Behavior of Toughened Zirconia Ceramics, Wear, 132, 15171 (1989). 3 W. Bundschuh and K.-H. Zum Gahr, Eect of Temperature on Friction and Oscillating Sliding Wear of Dense and Porous 3Y-TZP Zirconia Ceramics, Tribol. Int., 27, 97103 (1994). 4 A. Gangopadhyay, S. Jahanmir, and M. B. Peterson, Self-Lubricating Ceramic Matrix Composites; pp. 16399 in Friction and Wear of Ceramics, Edited by S. Jahanimr. Marcel Dekker, Inc., New York, 1994. 5 T. E. Fischer, M. P. Anderson, and S. Jahanmir, Inuence of Fracture Toughness on the Wear Resistance of Yttria-Doped Zirconium Oxide, J. Am. Ceram. Soc., 72 [2] 2527 (1989). 6 M. Woydt, J. Kadoori, K.-H. Habig, and H. Hausner, Unlubricated Sliding Behavior of Various Zirconia-Based Ceramic, J. Eur. Ceram. Soc., 7, 135 45 (1991).
1

4434

Journal of the American Ceramic SocietyVale et al.

Vol. 94, No. 12

7 H. Liang, T. E. Fischer, M. Nauer, and C. Carry, Eect of Grain Boundary Impurities on the Mechanical and Tribological Properties of Zirconia Surfaces, J. Am. Ceram. Soc., 76 [2] 3259 (1993). 8 S. W. Lee, S. M. Hsu, and M. C. Shen, Ceramic Wear Maps: Zirconia, J. Am. Ceram. Soc., 76 [8] 193747 (1993). 9 W. M. Rainforth, The Wear Behavior of Oxide Ceramics-A Review, J. Mater. Sci., 39, 670521 (2004). 10 R. Khanna and B. Basu, Sliding Wear Properties of Self-Mated YttriaStabilized Tetragonal Zirconia Ceramics in Cryogenic Environment, J. Am. Ceram. Soc., 90 [8] 252534 (2007). 11 H. G. Scott, Friction and Wear of Zirconia at Very low Sliding Speed; pp. 812 in Wear of Materials, Edited by K. C. Lumeda. American Society of Mechanical Engineer, New York, 1985. 12 S. Ran, L. Winnubst, D. H. A. Blank, H. R. Pasaribu, J. W. Sloetjes, and D. J. Schipper, Dry-Sliding Self-Lubricating Ceramics: CuO Doped 3YTZP, Wear, 267, 1696701 (2009). 13 X. Qunji and L. Huiwen, The Eect of Solid Lubricants on the Tribological Behavior of Zirconia at High Temperatures up to 600C, J. Phys. D: Appl. Phys., 30, 196571 (1997). 14 J. H. Ouyang, Y. F. Li, Y. M. Wang, Y. Zhou, T. Murakami, and S. Sasaki, Microstructure and Tribological Properties of ZrO2(Y2O3) Matrix Composites Doped With Dierent Solid Lubricants From Room Temperature to 800 C, Wear, 267, 135360 (2009). 15 J. Song, M. Vale, M. de Rooij, and D. J. Schipper, A Mechanical Model for Surface Layer Formation on Self-Lubricating Ceramic Composites, Wear, 268, 10729 (2010). 16 H. E. Sliney, Solid Lubricant Materials for High Temperatures-a Review, Tribol. Int., 15, 30315 (1982). 17 J. S. Zabinski, J. H. Sanders, J. Nainaparampil, and S. V. Prasad, Lubrication Using a Microstructurally Engineered Oxide: Performance and Mechanisms, Tribol. Lett., 8, 10316 (2000). 18 M. Goto, A. Kasahara, T. Oishi, Y. Konishi, and M. Tosa, Lubricative Coatings of Copper Oxide for Aerospace Applications, J. Appl. Phys., 94, 21104 (2003). 19 H. R. Pasaribu, J. W. Sloetjes, and D. J. Schipper, Friction Reduction by Adding Copper Oxide into Alumina and Zirconia Ceramics, Wear, 255, 699 707 (2003). 20 J. Song, M. Vale, M. de Rooij, D. J. Schipper, and L. Winnubst, The Eect of an Alumina Counterface on Friction Reduction of CuO/3Y-TZP at Room Temperature, Wear, in press (2011), doi: 10.1016/j.wear.2011.08.016. 21 B. Prakash and J. Celis, The Lubricity of Oxides Revised Based on a Polarisability Approach, Tribol. Lett., 27, 10512 (2007). 22 S. Ran, L. Winnubst, D. H. A. Blank, H. R. Pasaribu, J. W. Sloetjes, and D. J. Schipper, Eect of Microstructure on the Tribological and Mechanical Properties of CuO-Doped 3Y-TZP Ceramics, J. Am. Ceram. Soc., 90 [9] 2747 52 (2007). 23 G. Anstis, P. Chantikul, B. Lawn, and D. Marshall, A Critical Evaluation of Indentation Techniques for Measuring Fracture Toughness: I, Direct Crack Measurements, J. Am. Ceram. Soc., 64 [9] 5338 (1981). 24 H. Toraya, M. Yoshimura, and S. Somiya, Calibration Curve for Quantitative Analysis of the Monoclinic-Tetragonal ZrO2 System by X-ray Diraction, J. Am. Ceram. Soc., 67 [6] C11921 (1984). 25 X. Dong, S. Jahanmir, and S. M. Hsu, Tribological Characteristics of a-Alumina at Elevated Temperatures, J. Am. Ceram. Soc., 74 [5] 103644 (1991).

26 X. Dong and S. Jahanmir, Wear Transition Diagram for Silicon Nitride, Wear, 165, 16980 (1993). 27 J. F. Moulder, W. F. Stickle, P. E. Sobol, and K. D. Bomben, Handbook of X-ray Photoelectron Spectroscopy, Perkin-Elmer Corporation, Eden Prairie, MN, 1992. 28 J. Ghijsen, L. H. Tjeng, J. van Elp, H. Eskes, J. Westerink, G. A. Sawatzky, and M. T. Czyzyk, Electronic Structure of Cu2O and CuO, Phys. Rev. B, 38, 1132230 (1988). 29 F. Toschi, C. Melandri, P. Pinasco, E. Roncari, S. Guicciardi, and G. de Portu, Inuence of Residual Stresses on the Wear Behavior of Alumina/AluminaZirconia Laminated Composites, J. Am. Ceram. Soc., 86 [9] 154753 (2003). 30 J. R. Seidensticker and M. J. Mayo, Thermal Analysis of 3-mol%-YttriaStabilized Tetragonal Zirconia Powder Doped with Copper Oxide, J. Am. Ceram. Soc., 79 [2] 4016 (1996). 31 L. Lemaire, S. M. Scholz, P. Bowen, J. Dutta, H. Hofmeister, and H. Hofmann, Eect of CuO Additives on the Reversibility of Zirconia Crystalline Phase Transitions, J. Mater. Sci., 34, 220715 (1999). 32 L. Winnubst, S. Ran, E. A. Speets, and D. H. A. Blank, Analysis of Reactions During Sintering of CuO-Doped 3Y-TZP Nano-Powder Composites, J. Eur. Ceram. Soc., 29, 254957 (2009). 33 S. Ran, L. Winnubst, W. Wiratha, and D. H. A. Blank, Sintering Behavior of 0.8 mol%-CuO-Doped 3Y-TZP Ceramics, J. Am. Ceram. Soc., 89 [1] 1515 (2006). 34 B. Basu, R. G. Vitchev, J. Vleugels, J. P. Celis, and O. Van Der Biest, Inuence of Humidity on the Fretting Wear of Self-Mated Tetragonal Zirconia Ceramics, Acta Mater., 48 [10] 246171 (2000). 35 H. S. Kong and M. F. Ashby, Friction-Heating Maps and Their Application, MRS Bull., 16, 418 (1991). 36 N. P. Bansal and D. Zhu, Thermal Conductivity of ZirconiaAlumina Composites, Ceram. Int., 31, 9116 (2005). 37 A. M. Limarga and D. R. Clarke, The Grain Size and Temperature Dependence of the Thermal Conductivity of Polycrystalline, Tetragonal Yttria-Stabilized Zirconia, Appl. Phys. Lett., 98, 211906 (2011). 38 S. Schreck and M. Rohde, Preparation and Characterization of Ceramics Laser Alloyed with WO3 and CuO Nanopowders, Microelectron. J., 40, 681 6 (2009). 39 O. O. Ajayi, A. Erdemir, R. H. Lee, and F. A. Nichols, Sliding Wear of Silicon Carbide-Titanium Boride Ceramic-Matrix Composite, J. Am. Ceram. Soc., 76 [2] 5117 (1993). 40 Y. Jin, K. Kato, and N. Umehara, Tribological Properties of Self-Lubricating CMC/Al2O3 Pairs at High Temperature in air, Tribo. Lett., 4, 24350 (1998). 41 H. E. Sliney, T. P. Jacobson, D. Deadmore, and K. Miyoshi, Tribology of Selected Ceramics at Temperatures to 900C, Ceram. Eng. Soc. Proc., 7, 103951 (1986). 42 A. G. Atkins and D. Tabor, Hardness and Deformation Properties of Solids at Very High Temperatures, Proc. R. Soc. London Ser. A, 292, 44159 (1966). 43 G. Vagnard and J. Washburn, Plastic Deformation of Cuprous Oxide, J. Am. Ceram. Soc., 5 [2] 8894 (1968). h

Das könnte Ihnen auch gefallen