Anda di halaman 1dari 8

Surface and Coatings Technology 146 147 (2001) 201208

Structureperformance relations in nanocomposite coatings


Jorg Patscheidera,*, Thomas Zehndera, Matthieu Diserensb
a

Department of Surface and Joining Technology, EMPA Dubendorf, CH-8600 Dubendorf, Switzerland b Institut de Physique Appliquee, EPFL, 1015 Lausanne, Switzerland

Abstract Properties of hard nanocomposite coatings, especially hardness, can be explained by their nanostructure. Hardness maxima are found for different nanocrystallineyamorphous materials deposited by different techniques at typically 20% of the amorphous phase. A zone model is proposed which correlates the hardness to the relative phase content. The hardness of nanocomposite coatings peaks at the common minimum of the grain size of the crystalline phase and the grain separation. For an adequate description of the performance of a coating, the thermal stability, oxidation behavior and frictional behavior should be included in addition to hardness. In a friction situation involving at least two friction partners, the overall behavior of the system is determined by many-body interactions. While thermal stability and oxidation properties as inherent material properties can be directly linked to the nanostructure of the coating, the frictional behavior of a coating cannot be generalized independent of the friction conditions. 2001 Elsevier Science B.V. All rights reserved.
Keywords: Hard nanocomposite coatings; Nanostructure; Hardness maxima

1. Introduction Wear-protective hard coatings have served for more than 30 years to prevent the untimely consumption of surfaces exposed to wear. Since their introduction to the market in the early 1980s, single-phased materials like TiN, TiC and CrN have been the most widely used materials to prevent surface damage mainly by abrasion. Next generation materials showed improved properties like hardness up to 30 GPa, lower adhesion of the surface to the friction partner, better oxidation resistance (up to 8008C in air) and many others; these improvements were mainly achieved by ternary materials like TiAlN w1,2x. Multilayered heteroepitactic coatings, consisting of alternating layers of two hard phases (e.g. TiN with VN or NbN), show a hardness enhancement by more than a factor of two, when the thickness of the individual layers was of the order of several nanometers. Differences in the shear moduli of the phases and low* Corresponding author. Tel.: q41-1-823-43-65; fax: q41-1-82340-34. E-mail address: joerg.patscheider@empa.ch (J. Patscheider). 0257-8972/01/$ - see front matter PII: S 0 2 5 7 - 8 9 7 2 0 1 . 0 1 3 8 9 - 5

energy interfaces were identified as a prerequisite for enhanced hardness of these material combinations w3 5 x. Analogous to the nanoscaled multilayer coatings, it is possible to deposit thin films, which show similar hardness enhancement by combining two materials isotropically. Such coatings consist, as is the case for multilayers, of two non-miscible phases with distinct phase boundaries. In order to achieve hardness enhancement effects similar to those found in multilayers, the dimensions between two phase boundaries, i.e. the average size of the crystallites, has to be of the same size as the thickness of a single layer in a multilayer system, which is of the order of 5 nm. Because such materials are composed of two different materials with grain size between 3 and 10 nm, they are called nanocomposites. An extended review article on the existing literature on nanocomposite hard coatings can be found in Veprek w6x. Nanocomposites can either consist of two nanocrystalline materials as it is the case for ZrNyCu w7x, TiNycBN w8x, TiB2 yTiN w9x and similar combinations or of a nanocrystalline material in combination with an amorphous phase. Examples for

2001 Elsevier Science B.V. All rights reserved.

202

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

the latter case are nc-TiNya-Si3 N4 , nc-TiCya-C:H (both discussed later in this paper), nc-TiCya-C w10x and several others. Hardness values of such coatings are reported to be as high as 105 GPa w11x. Despite the importance of hardness as a key parameter for wearresistant coatings, other properties decisive for practical applications like frictional behavior and thermal stability of nanocomposite coatings have been discussed less extensively. In this paper we will try to evaluate the importance of high hardness with respect to other important properties of hard nanocomposite coatings and to correlate observed properties with nanostructural aspects. 2. Hard nanocomposite coatings: nc-TiNya-Si3 N4 and nc-TiCya-C:H Following the concept of incorporating stable oxideforming elements (Al, Si, Hf, Cr, Zr, Nb) into TiN w1,2x, efforts were undertaken to codeposit silicon and titanium nitride. In contrast to TiAlN, TiZrN and other single-phased hard materials, silicon cannot be substitutionally built in the lattice of TiN. In accordance with the TiSiN phase diagram which does not present any stable ternary phase under equilibrium conditions w12x, two-phase TiNySi3N4 coatings form when silicon is added during deposition of TiN. Despite some indications that an unidentified compound is formed under conditions of low ion bombardment w13x, no indications of incorporation of silicon into the lattice of TiN could be found w14x. The first reported coatings consisting of TiSiN have been produced by chemical vapor deposition (CVD) in 1982 by Hirai et al. w15x. Nicolet produced TiSiN coatings by reactive sputtering from TiSix-targets as electrically conductive diffusion barriers for copper metallization w16,17x. The first successful attempts to improve the hardness of TiN by addition of silicon have been reported by Li et al. w18x and Veprek et al. w1922x. They deposited TiNSi3N4 films by plasma-enhanced chemical vapor deposition (PACVD) at deposition temperatures of 5506008C. Li used TiCl4, SiCl4 and H2 in his process, while Vepreks group used SiH4 instead of SiCl4. These films showed unusually high hardness with a maximum of approximately 50 GPa at 7 atomic percent (at.%) silicon in the TiSi N film. These coatings were identified as nanocomposites consisting of TiN nanocrystallites of approximately 47 nm (nc-TiN) surrounded by an amorphous Si3N4 (a-Si3N4) matrix. The early experiments for deposition of nc-TiNyaSi3N4 by PACVD used SiCl4, SiH4, TiCl4 and H2. These gases pose a series of problems for manufacturing (corrosion of film and of equipment, fire and environmental hazards); in order to avoid these problems, todays deposition technologies are all based on PVD techniques w7,13,23,24x. Unbalanced magnetron sputter-

Fig. 1. Comparison of hardness values of nc-TiNya-C:H nanocomposites prepared by different research groups w8,13,23x.

ing (UBM), a magnet arrangement in a conventional planar magnetron, where the outer magnets are stronger than the central ones, is a technique which provides sufficiently high degrees of dissociation and ionization at the substrate w25x. Superimposing a glow discharge on the substrates placed on the RF-powered sample holder increases the surface mobility of adsorbed species on the growing film and thus provides a tool for depositing dense films at relatively low substrate temperatures. The two nanocomposites described in this paper have been deposited by this technique; details of the deposition process are given by Patscheider and coworkers w23,26x. 2.1. Hardness and morphology A comparison of nc-TiNya-Si3N4 nanocomposites described by different groups w8,23,27x shows that the hardness is a distinct function of the silicon concentration in the films (see Fig. 1). Although the different authors used different deposition techniques (PACVD and reactive UBM-PVD), all of them observed a welldefined hardness maximum which peaks between 15 and 25 at.% of the amorphous phase. The hardness lies between 40 and 60 GPa; mean crystallite sizes of 510 nm are measured at maximum hardness. However, much higher hardness values have recently been reported for materials deposited by PACVD, where hardness values exceeding the one of diamond are postulated w11,28x. 2.2. Zone model for nanohardness nanostructure relations The ratio of the amorphous to the nanocrystalline phase has a strong influence on the orientation of the

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

203

Fig. 2. (a) Preferred orientation of the TiN nanocrystallites in nc-TiNy a-Si3N4 from X-ray diffraction (I111yI200 ) and mean crystallite size perpendicular to the surface. (b) The comparison shows three zones with different characteristics (see text and Fig. 3).

nanocrystals in the coatings (cf. Fig. 2a). While siliconfree TiN grows with a pronounced preferred orientation in N111M, as it is known for PVD TiN, only minor amounts of silicon (13 at.%) in the films lead to a loss of the N111M orientation and randomly oriented crystallites are deposited. Coatings with only 3 at.% of silicon show no preferred orientation. A similar development of the mean size of the TiN grains, as measured perpendicular to the surface, is observed as the silicon content is increased. The mean grain size (perpendicular to the surface) of TiN deposited without silicon is more than 100 nm. Codeposition of a-Si3N4 causes a marked decrease of the crystallite size, which reaches values of less than 10 nm at 5 at.% of silicon in the deposited coating. Fig. 2a displays the preferred orientation of the TiN crystallites (intensity ratio of the w111x and w200x X-ray diffraction peaks) and the mean grain size vs. the silicon content in the films. A comparison of the hardness (cf. Fig. 2b) enables us to subdivide the full compositional range between 0% silicon nitride (i.e. only TiN) and silicon nitride as the only phase (i.e. no TiN) into three zones A, B and C. The first zone, marked by A in Fig. 2, is confined by the sudden loss of the N111M orientation and is bordered between 0 and approximately 3 at.% silicon. This zone is characterized by a sharp increase of the hardness and a concomitant decrease of the mean crystallite size from well above 100 nm to values below 10 nm at 3 at.% silicon.

Silicon-free titanium nitride is not hindered in its growth by amorphous silicon nitride and forms therefore large elongated crystallites. Addition of minor amounts of silicon nitride leads to an encapsulation of the growing TiN crystallites and thus hinders their further growth. The sudden drop of the orientation supports this assumption since an isotropic growth of TiN (i.e. no preferred orientation) is observed. At 1 at.% silicon in the coating and assuming a TiN crystallite shape of 100=10=10 nm, the averaged coverage of such a crystallite would be approximately half a monolayer of Si3N4. Even if silicon nitride would be uniformly distributed over such a crystals surface, full coverage cannot be reached and therefore TiN growth can continue. At approximately 34 at.% silicon in the layers the entire crystallite surfaces are statistically covered with a monolayer of Si3N4. Due to the finite surface energy of Si3N4, a continuous monolayer is an energetically unfavorable situation, and therefore coverage will be incomplete and growth of TiN crystallites, although limited by the partially covering Si3N4, can occur. These changes in the nanostructure have a pronounced influence on the hardness of the films as evidenced by the steep rise of the nanohardness values within this zone A. When comparing the immediate rise of the hardness with the paralleled vanishing of the preferred orientation, it is tempting to assume that the development of equiaxed grains is an important factor for obtaining enhanced hardness. It is, however, not the only reason, as is be evidenced from zone B. The following regime on the concentration scale of silicon between 3 and 10 at.% is labeled with B. Above 34 at.% Si, equiaxed grains of TiN with average size between 5 and 10 nm coexist with amorphous silicon nitride which completes coverage of the crystallites, as the silicon content is raised to approximately 10 at.%. Nanohardness values reach a maximum of 40 GPa at approximately 10 at.% silicon. At this concentration the mean grain separation by the amorphous phase amounts to approximately 0.4 nm which corresponds to four chemical bond lengths in Si3N4. This thin layer of Si3N4 is apparently sufficient to further promote the hardness of the composite. This is probably the percolation threshold, i.e. the point where a full interpenetration of the two phases is reached w29x. Although no change of the grain orientation and only insignificant decrease of the grain size is observed, a prominent hardness loss occurs when the silicon content is raised above 10 at.%. This marks the starting point of the third zone C in the hardness curve. Once the percolation threshold is reached and the mean grain separation by the amorphous phase exceeds a certain value, which amounts to approximately 0.5 nm in the case of this type of nanocomposites, the interaction of the two phases observed in zone B is lost and the hardness of the nanocomposite is determined by the

204

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

properties of the amorphous phase alone. This means that the hardness of the nanocomposite approaches the one of Si3N4. The zone model sketched above can be illustrated in a morphology scheme of the grain size and shape as a function of the silicon contribution, as it is displayed in Fig. 3. It depicts the large elongated grains at very low silicon contents and the successive loss of preferred orientation in zone A, the full interpenetration of both phases in zone B and the isolated TiN nanocrystals in zone C. It was shown for a series of other nanocrystalliney amorphous nanocomposite coatings that enhanced hardness values can be obtained, if the proper ratio of the nanocrystalline phase to the amorphous one is maintained w22,26,29,30x. This is e.g. the case for nc-TiCy a-C:H w25x which was deposited by reactive UBM sputtering following the principles developed for ncTiNya-Si3N4 w23,31x. Hardness enhancement by a factor of two, leading to a hardness maximum of 35 GPa, could be observed at 20% of the amorphous phase. Raman measurements which are displayed in Fig. 4a, show that the amorphous phase in nc-TiCya-C:H is composed of a-C:H; it is detectable from approximately 20 at.% up to TiC-free a-C:H (see Fig. 4b). Curve fitting following established procedures w32,33x indicated structural changes towards less extended graphite islands between 40 and 80% of the amorphous phase as evidenced by the ratio of the peaks from disordered graphite (D peak) and the crystalline graphite (G peak) (see Fig. 4c). The nature of the amorphous phase changes slightly as the mean grain separation shrinks to values below 1 nm. The Raman data from Fig. 4c

suggest that the nature of the amorphous hydrogenated carbon between the grain boundaries of TiC becomes less disordered as the mean grain separation decreases. Fig. 5 displays the hardness of nc-TiCya-C:H along with the grain size and the mean distance between the crystallites. This comparison shows that for maximum hardness two conditions have to be fulfilled: the nanocrystals are approximately 5 nm in size and the mean distance has to be very small. The calculations, using the crystallite size and the phase composition, show that the mean grain separation is approximately 0.4 nm. If the separation is too large (e.g. at 40% of the amorphous phase, the hardness is mainly governed by the properties of the amorphous network. In the absence of the amorphous phase, the material behaves like a polycrystalline material and low (i.e. regular) hardness is obtained. This is again an example how the hardness of a nanocomposite is controlled by its nanostructure. 3. Performance of nanocomposite hard coatings The application potential of a coating for a certain use is given by its properties. Modern hard coatings have to fulfil more than one property in order to replace older coatings that are well established in the market. Beside the necessary prerequisite of having at least the hardness of TiN, oxidation resistance, enhanced thermal stability or improved tribological performance are qualities that are required for new coatings. Since high hardness alone is not a property that determines the performance of a coating w34x, it is necessary to know the behavior in situations close to applications. Nanocomposite hard coatings provide, apart from their high

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

205

Fig. 4. (a) Normalized Raman spectra of nc-TiCya-C:H at different a-C:H contents. D2 (1130 cmy1 ) is assigned to amorphous sp3, D1 (1350 cmy1) to polycrystalline (disordered) graphite, G2 (1490 cmy1 ) to amorphous sp2 and G1 (1585 cmy1) to the E2g mode of graphite w32x. (b) Total Raman intensity vs. a-C:H phase content. (c) Intensity ratio I(D1qD2)yI(G1qG2) vs. a-C:H phase content (see text).

hardness at the typical composition, other qualities, which make them promising candidates for industrial applications. Two of the most important ones are the thermal stability and the properties, which determine their tribological behavior. Both of them shall be discussed in the following section. 3.1. Tribological behavior Enhanced performance of a wear-protective coating can manifest itself, e.g. as reduced energy losses in a sliding process due to a self-lubricating coating w35x, as prolonged lifetime of a coated cutting tool in a chipping process w36x or by enabling mechanical movements which would not be possible without proper coatings w37x. In contrast to hardness, the tribological behavior of a coating is not a materials property. The tribological performance of a coating can be meaningfully described only for a given set of conditions, since the frictional behavior is strongly governed by the properties of the friction partners and the conditions (temperature, atmosphere, pressure, type of movement, load, etc.) of the friction. Friction takes place when two solids with a relative motion against each other come into contact. Depending on the environmental, geometrical and mechanical con-

ditions, different behavior will result. Any experiment to assess the tribological behavior of surfaces will therefore depict only a part of the reality. A widely used experimental setup to measure friction coefficients and wear rates is the pin-on-disk test. The friction data given in the following are obtained from pin-on-disk tests in controlled humidity of 60% RH, measured against a 100Cr6 steel ball (diameter 6 mm, loaded at 5 N) at a rotation speed of 10 cmys and a track diameter of 16 mm. Data for the wear rate are taken after 720 m of traveled distance. 3.1.1. nc-TiNya-Si3N4 Fig. 6 shows the friction coefficient and wear rate of nc-TiNya-Si3N4 coatings with maximum hardness. It can clearly be seen that both the friction coefficient as well as the wear rate of the coating show no dependence on the composition, which is in sharp contrast to the hardness dependence on the composition. The friction coefficient of nc-TiNya-Si3 N4 , remaining constant at 0.9 over the whole compositional range, is even worse than the one of TiN, which amounts to 0.6. The wear rate of the coatings under these conditions, as calculated from the depth of the wear track, is correspondingly higher than that of TiN. This is a very undesired finding for a

206

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

Fig. 6. Friction coefficient and wear rate of nc-TiNya-Si3 N4 against 100Cr6 from pin-on-disk measurements (conditions see text).

Fig. 5. Comparison of nanohardness, mean grain size and mean grain separation in nc-TiCya-C:H. The hardness maximum occurs at the common minimum of grain size and grain separation.

appearance of the amorphous phase leads to a drop of the wear rate by a factor of approximately 100. This change of the wear rate is paralleled by the hardness maximum. An increase of the amorphous phase contribution lowers the wear rate even more. This is explained by the formation of a transfer layer under contact

coating with very high hardness (and therefore good potential for industrial applications). However, in tribological situations with a liquid phase, where a direct contact of coating and counterpart is not given, nc-TiNya-Si3N4 showed a much improved wear behavior as compared to TiN w23x. In a ball cratering test (CSEM Calowear) carried out with hard abrasive suspensions (alumina slurry and diamond grits in methanol), only 1 at.% of silicon in layers lowered the wear rate by a factor of 8. Coatings with higher silicon concentrations (cSi)5 at.%) had still lower wear rates, though not as low as those with very low silicon concentrations. The wear behavior in this tribological situation seems to be linked to the grain orientation and, at silicon concentrations above 5 at.%, governed by the phase contribution of Si3N4. Again, no correlation with the hardness of the coatings is observed, although the nanostructure of the coatings seem to influence the behavior under this set of tribological conditions. 3.1.2. nc-TiCya-C:H The wear rate of the 100Cr6 steel ball against ncTiCya-C:H coatings as a function of their phase composition, along with the hardness, is displayed in Fig. 7. In contrast to nc-TiNya-Si3 N4 , a marked influence of the composition on the wear rate (and also of the friction coefficient w26x) is observed. While TiC coatings (i.e. no a-C:H) exhibit a relatively high wear rate,

Fig. 7. Comparison of nanohardness of nc-TiCya-C:H and wear rate of 100Cr6 steel ball against nc-TiCya-C:H (conditions see text). The wear rate is steered by the amount of a-C:H that can be transformed into a self-lubricating transfer film.

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208

207

pressure. Since a-C:H can transform into graphite-like material with very low shear strength, an increase of the a-C:H fraction leads to enhanced formation of this selflubricating transfer layer resulting in a low friction coefficient and low wear w38x. On the other hand, when no a-C:H is available for transformation, the friction is controlled by the properties of TiC and hence the friction coefficient and the wear rate rise in a discontinuous way. At the composition of the hardness maximum (20% a-C:H) a friction coefficient of 0.3 is measured. Because of the ability of a-C:H to transform into a lubricating material (in contrast to a-Si3 N4 ) for a certain compositional range (20100%), nc-TiCya-C:H is an example for a case where the composition and the nanostructure strongly influence the friction properties. 3.2. Thermal stability 3.2.1. nc-TiNya-Si3N4 The limited oxidation resistance of TiN was one of the incentives to improve this material by the addition of elements forming stable and sufficiently hard oxides w1,2x. Silicon as an element, which forms highly hard and stable oxides, was expected to promote the oxidation stability of nc-TiNya-Si3 N4 . An enhanced resistance against oxidation was reported already in one of the first publications on hard nanocomposites w19x. The oxidation of nc-TiNy-a-Si3 N4 proceeds via a mechanism with at least two different steps. A comparison of the oxide thickness of coatings with different amounts of Si3N4 at temperatures between 600 and 10008C in air showed that coatings with large silicon contents oxidize with a high apparent activation energy (resulting in low rates) up to a certain temperature above which accelerated oxidation with an activation energy typical for TiN takes place. This temperature, at which the change in mechanism occurs, is a function of the silicon content in the coating w31x. Amorphous Si3N4 as an efficient diffusion barrier for oxygen and other elements w16,17x causes low rates of formation of SiO2, which itself is also a good diffusion barrier for oxygen. The thicker the aSi3N4 layer is, the longer it takes oxygen to reach the encapsulated TiN crystallites. Once oxidation of TiN commences, crack opening of the SiOx scale occurs due to the increased molar volume of TiO2 as compared to the one of TiN. Through theses cracks access of oxygen to the reactive zone is facilitated and accelerated oxidation of TiN is observed. This stepwise oxidation mechanism results in low oxidation rates at high silicon contents and almost same rates as for TiN in the case of low silicon concentrations. Nevertheless, at the composition of the hardness maximum, the oxidation proceeds approximately 20 times slower as compared to TiN. The thin layers of a-Si3N4 surrounding the nanocrystallites are efficient diffusion barriers also for titanium

w39x and nitrogen w40x. A consequence thereof is the stability against recrystallization, which requires unhindered diffusion of Ti and N atoms. Due to the restricted self-diffusion, the grain size with the surrounding amorphous phase and hence the hardness is maintained up to 9508C for 1 h w14x or up to 11008C for 30 min w8x. 3.2.2. nc-TiCya-C:H Hydrogenated amorphous carbon (a-C:H) is a class of amorphous materials with limited temperature stability. When exposed to temperatures above approximately 4008C the material starts to liberate chemically bonded hydrogen and formation of disordered graphitic material with finite size of the graphitic planes begins. This change can be followed using Raman scattering which has proven to be a valid method to characterize the thermal degradation of a-C:H w33x. It is therefore expected that nc-TiCya-C:H, at least for high fractions of aC:H, behaves in a similar way. A structural change of the amorphous phase above 4008C has been observed using Raman spectroscopy in terms of I(D)yI(G) as it was found in annealing experiments of TiC-free a-C:H w32x. 4. Conclusions Many of the technologically exploitable properties of nanocomposite hard coatings like hardness and thermal stability are directly related to their nanostructure. The hardness of a single-phase material can be understood from very basic material properties like bond strength, bond energy density and shear modulus w41x. The hardness of nanocomposites, however, is related to a complex interplay of size of the nanocrystallites, their orientation and the separation of the crystallites by the amorphous phase. High hardness is obtained when the crystallite size is approximately 5 nm and the fully percolated amorphous phase statistically separates the crystallites by approximately 0.5 nm. The hardness is directly linked to the nanostructure of the nanocomposite. In contrast to hardness and other materials properties, the tribological performance of a coating is the result of complex interactions of various parameters. For example, nc-TiCya-C:H has low friction because a solid lubricant forms under load. nc-TiNya-Si3 N4 cannot undergo a similar transformation (in unlubricated situations) and shows the tribological behavior known from TiN and Si3N4. It was shown that nc-TiNya-Si3N4 performs well in a situation involving a liquid and that the same coating fails if the conditions are changed (e.g. no liquid partner) w14x. Tribological experiments show clearly that hardness alone cannot be taken as a sufficient parameter to describe the performance of a coating in a given situation. Despite the obvious need for tribological characterization of a protective coating in

208

J. Patscheider et al. / Surface and Coatings Technology 146 147 (2001) 201208 w11x A. Niederhofer, P. Nesladek, H.-D. Mannling, K. Moto, S. Veprek, M. Jilek, Surf. Coat. Technol. 120-121 (1999) 173. w12x P. Rogl, J.C. Schuster, Phase Diagrams of Ternary Boron Nitride and Silicon Nitride Systems, ASM International, Materials Park (OH), 1992. w13x F. Vaz, L. Rebouta, P. Goudeau et al., Surf. Coat. Technol. 133134 (2000) 307. w14x M. Diserens, PhD thesis, Swiss Fed. Inst. Technol., Lausanne (2000). w15x T. Hirai, S. Hayashi, J. Mater. Sci. 17 (1982) 1320. w16x X. Sun, J.S. Reid, E. Kolawa, M.-A. Nicolet, J. Appl. Phys. 81 (2) (1997) 656. w17x Y. Tsuji, S.M. Gasser, E. Kolawa, M.-A. Nicolet, Thin Solid Films 350 (1999) 1. w18x L. Shizhi, S. Yulong, P. Hongrui, Plasma Chem. Plasma Process. 12 (3) (1992) 287. w19x S. Veprek, S. Reiprich, L. Shizhi, Appl. Phys. Lett. 66 (20) (1995) 2640. w20x S. Veprek, Pure Appl. Chem. 68 (5) (1996) 1023. w21x S. Veprek, S. Reiprich, Thin Solid Films 268 (1995) 64. w22x S. Veprek, M. Haussmann, L. Shizhi, Proc. Electrochem. Soc. 96 (5) (1996) 619. w23x M. Diserens, J. Patscheider, F. Levy, Surf. Coat. Technol. 108 109 (1998) 241. w24x A. Niederhofer, P. Nesladek, H.D. Mannling, K. Moto, S. Veprek, M. Jilek, Surf. Coat. Technol. 121 (1999) 173. w25x B. Window, N. Savvides, J. Vac. Sci. Technol. A 4y2 (1986) 196. w26x T. Zehnder, J. Patscheider, Surf. Coat. Technol. 133-134 (2000) 13. w27x F. Vaz., L.M. Rebouta, P. Goudeau et al., Surf. Coat. Technol. 133-134 (2000) 307. w28x P. Nesladek, S. Veprek, Phys. Stat. Solidi A 177 (1) (2000) 53. w29x S. Veprek, M. Haussmann, S. Reiprich, J. Vac. Sci. Technol. A 14 (1996) 46. w30x A.A. Voevodin, J.M. Schneider, C. Rebholz, A. Matthews, Tribol. Int. 29y7 (1996) 559. w31x M. Diserens, J. Patscheider, F. Levy, Surf. Coat. Technol. 120 121 (1999) 158165. w32x U. Muller, R. Hauert, Mater. Res. Soc. Symp. Proc. 434 (1996) 113118. w33x R.O. Dillon, J.A. Woolam, V. Katkanant, Phys. Rev. B 29y6 (1984) 3482. w34x A. Leyland, A. Matthews, Wear 246 (1-2) (2000) 1. w35x R. Gilmore, M.A. Baker, P.N. Gibson et al., Surf. Coat. Technol. 108-109 (1998) 345. w36x P.E. Hovsepian, W.-D. Munz, A. Medlock, G. Gregory, Surf. Coat. Technol. 133-134 (2000) 508. w37x A.A. Voevodin, J.S. Zabinski, Thin Solid Films 370 (2000) 223231. w38x A. Grill, Surf. Coat. Technol. 94-95 (1997) 507. w39x C.J. Bedell, H.E. Bishop, G. Dearnaley, J.E. Desport, H. Romary, J.M. Romary, Nucl. Instr. Methods Phys. Res. B 59-60 (1991) 245. w40x R.M.C. de Almeida, I.J.R. Baumvol, Phys. Rev. B 62 (24) (2000) R16255. w41x D.M. Teter, MRS Bull. 23y1 (1998) 22.

order to judge its application potential, it must be emphasized that tribological data are valid only for one given ensemble of tribological parameters. Extrapolation of friction data from one series of tribological experiments to another one can often not be made; delineation of causes and generalizations are therefore in most cases not possible. The high hardness of nc-TiNya-Si3 N4 , combined with its excellent stability against oxidation and recrystallization, makes it a valid candidate material for applications in unlubricated situations where high mechanical load at high temperature act on a workpiece (e.g. extrusion, die casting, molding). Its high friction coefficient against steel in the absence of a lubricant, however, may impose limits to applications where hard coatings with better frictional behavior are already on the market. If a hard coating is required that provides self-lubricating properties, nc-TiCya-C:H should be considered. The low wear rate against steel along with the good hardness makes the coating interesting for applications where low friction (and hence no excessively high thermal loads) at high mechanical loads occur, as it is the case in bearings where lubricants are undesired or not possible. Acknowledgements The authors are grateful for financial support by the presidents funds of EMPA and EPFL as well as by the Swiss Priority Program on Materials (PPM). References
w1x W.-D. Munz, J. Vac. Sci. Technol. A 4y6 (1986) 2717. w2x O. Knotek, M. Bohmer, T. Leyendecker, J. Vac. Sci. Technol. A 4y6 (1986) 2695. w3x S. Barnett, A. Madan, Phys. World, Jan. (1998) 45. w4x W.D. Sproul, Science 273 (1996) 889. w5x H. Ljungcrantz, C. Engstrom, L. Hultman et al., J. Vac. Sci. Technol. A 16 (5) (1998) 3104. w6x S. Veprek, J. Vac. Sci. Technol. A 17 (5) (1999) 2401. w7x J. Musil, P. Zeman, H. Hruby, P. Mayrhofer, Surf. Coat. Technol. 121 (1999) 179. w8x S. Veprek, P. Nesladek, A. Niederhofer, F. Glatz, M. Jilek, M. Sima, Surf. Coat. Technol. 109y1-3 (1998) 138. w9x C. Mitterer, P.H. Mayrhofer, M. Beschliesser et al., Surf. Coat. Technol. 121 (1999) 405. w10x A.A. Voevodin, J.S. Zabinski, Diamond Relat. Mater. 7 (1998) 463.

Anda mungkin juga menyukai