Sie sind auf Seite 1von 8

Surface & Coatings Technology 201 (2007) 7441 7448 www.elsevier.

com/locate/surfcoat

Effect of Ni sublayer thickness on sliding wear characteristics of electrodeposited Ni/Cu multilayer coatings
S.K. Ghosh a,, P.K. Limaye b , S. Bhattacharya c , N.L. Soni b , A.K. Grover a
c

Materials Processing Division, Bhabha Atomic Research Centre, Trombay, Mumbai-400 085, India Refuelling Technology Division, Bhabha Atomic Research Centre, Trombay, Mumbai-400 085, India Technical Physics and Prototype Engineering Division, Bhabha Atomic Research Centre, Trombay, Mumbai-400 085, India
b

Received 16 November 2006; accepted in revised form 6 February 2007 Available online 15 February 2007

Abstract The electrodeposition of Ni/Cu multilayer coatings with varied Ni sublayer thickness (tNi) ranging 1.811 nm, keeping Cu sublayer thickness (tCu) fixed at 1.97 nm, has been carried out from sulphate based electrolyte. Residual stress and crystal structure of the electrodeposited multilayers were characterized by X-ray diffraction (XRD) method. It was found that the stress level within the deposited multilayer decreases with increase in tNi. Coefficient of friction (COF) and sliding wear test of the deposited multilayer coatings have been carried against standard WC ball in ball-onplate geometry under dry condition. After wear test, the wear scar morphology was examined under laser surface profilometry and scanning electron microscopy (SEM) for calculation of wear volume and evaluation of wear mechanism respectively. The COF was found to be dependent on load and tNi. A typical sinusoidal variation of COF with sliding distance was noticed indicating stickslip behaviour under adhesiveabrasive wear process. The hardness of the multilayers was found to be dependent on wavelength of the modulation. Among multilayers, sample with tNi = 7.2 nm has shown the lowest wear rate owing to its high (H/E) ratio, high hardness and low residual stress. The calculated wear rate order was found to be quite well matching with the order of elastic strain to failure, i.e. (H/E) ratio of the multilayers. The multilayers with very high and low tNi showed poor wear behaviour due to very weak multilayer effect. 2007 Elsevier B.V. All rights reserved.
Keywords: Ni/Cu multilayers; Residual stress; XRD; Electrodeposition; Coefficient of friction;Wear; Tribology

1. Introduction Metallic multilayers with near atomic sublayer thickness have been under investigation for a couple of decades because of their attractive mechanical, magnetic and optical properties [14]. There are various applications where multilayer structure plays an important role. Especially, metallic magnetic multilayers (MMM's) and spin-valve structures already find many applications in magnetic heads and in magnetic rigid disk drives, head tape devices, and various components in microelectromechanical systems (MEMS) devices. However, in these device applications, tribological behaviour of multilayer mate-

Corresponding author. Tel.: +91 22 25591722; fax: +91 22 25505151/ 25505239. E-mail address: sghosh@barc.gov.in (S.K. Ghosh). 0257-8972/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.surfcoat.2007.02.014

rials is a matter of concern. Typical layers applied in these devices are as thin as 1 nm. The tribological properties of these layered structures directly depend on the layer properties, e.g., the layer thickness, surface roughness, coefficient of friction, hardness, stiffness and elastic modulus ratios of the layers to the substrate. And for quite a long time, it has been known that it was the hardness of surface and subsurface layers [5] which are responsible for observed wear behaviour. Due to these reasons, most of the earlier effort was put to synthesize structures with enhanced hardness for obtaining better wear resistance [610]. Among various possibilities, alternate nano-laminated structure was one of them. In these structures, the strengthening is due to (a) presence of interface between two layers with wide difference in elastic modulus acts as dislocation pinning site (Koehler type strengthening) [11]. (b) The other one is substructure strengthening mechanism, and the yield strength (y)

7442

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

Table 1 Lattice parameters, bilayer period and microhardness of the electrodeposited Ni/Cu multilayers ID S1 S2 S3 S4 Sample [Ni(1.8 nm)/Cu(1.97 nm)] [Ni(4.5 nm)/Cu(1.97 nm)] [Ni(7.2 nm)/Cu 1.97 nm)] [Ni(11 nm)/Cu 1.97 nm)] d-spacing from XRD (nm) 0.207 0.205 0.204 Stress free d-spacing (nm) 0.2072 0.2052 0.2054 Bilayer period from EQCM (nm) 3.77 6.47 9.17 Bilayer period from XRD (nm) 4.56 6.48 9.48 Microhard-ness (KHN25gf) 318 353 390 340

depends on sublayer thickness as y p, where p is the exponent and the mechanism changes depending upon its values; when p = 1/2, the relation becomes y 1/2 which is nothing but HallPetch relationship (HallPetch strengthening) [12,13]. And if p = 1, the equation reduces to y 1, this is nothing but Orowan type strengthening and if p 1/2, it becomes modified HallPetch strengthening. Basically, this kind of mechanism depends on sublayer thickness, grain size within sublayers and dislocation pile-up. In fact, it was found that significant improvement in elastic modulus and hardness could be achieved for multilayer coatings with suitable modulation wavelength [1418]. However, by critical examination of previous results, Leyland and Matthews [19,20] recently illustrated that the ratio of hardness (H) to elastic modulus (E), H/E i.e. elastic strain to failure and fracture toughness are more appropriate factors in determining the tribological properties of metallic and ceramic nanocomposite coatings. Higher magnitude of H/E ratio would lead to better wear resistance of the multilayers coatings and this can be achieved either y increasing hardness (H) or by lowering elastic modulus (E) and the latter can be as low and as very close to substrate so that delamination across coating/substrate is avoided. In one of our recent study on Ni/Cu multilayers [21], it was noticed that not only H/E ratio, the residual stress within the deposited structure is also an important factor in determining its wear behaviour. A few studies are available in the literature regarding the tribogical properties of the Ni/Cu multilayers [2225]. In these studies, Ni/Cu multilayers deposited from sulphamate bath with thickness ranging from 3.5 to 100 nm were investigated keeping both sublayer thickness identical. It was found that the wear rate of Ni/Cu multilayer was lower than pure Cu and Ni coatings. Previous investigator have used steel ball or cylinder as counter body for sliding wear studies which led to the difficulty in wear volume measurement. To avoid these problems, we reported [21] that use of tungsten carbide (WC) ball in sliding geometry as mating body could be a better alternative. We have also shown how coefficient of friction, wear rate and mechanism of wear changes with tCu and concentrated mainly on Ni/Cu multilayers with sublayer thickness 11 nm. In the present study, sulphate based electrolyte has been used for the deposition of Ni/Cu multilayers coatings to avoid sulphur contamination. Residual stress of the Ni/Cu multilayer coatings has been investigated in the light of tNi/tCu ratio. The main objective was to study the effect of nickel layer thickness ( 11 nm) on wear and frictional behaviour. After wear test, the wear scar morphology was examined under SEM to know the wear mechanism and correlate with the observed results.

2. Experimental [Ni(x nm)/Cu(1.97 nm)]n multilayers, where n equals to total number of each layers, were deposited on polished stainless steel (SS) substrates from a single solution bath as mentioned in our earlier work [21]. The total thickness for Ni/Cu multilayer was kept at 10 m. Four films with multilayer structure of [Ni(tNi nm)/Cu(1.97 nm)] were studied with tNi as 2, 5, 8 and 11 nm and are denoted as S1, S2, S3 and S4 respectively as mentioned in Table 1. Deposition was carried out galvanostatically and pulse current was supplied from AUTOLAB PGSTAT30 potentiostat/galvanostat. The deposition current density was 0.2 mA cm 2 for copper and potentiostat/galvanostat. The deposition current density was 20 mA cm 2 for nickel. The typical sublayer thickness has been determined using electrochemical quartz crystal microbalance (EQCM) of Maxtek PM-710 attached with PGSTAT30. The other details of sample preparation were given elsewhere [21]. X-ray diffraction (XRD) of these deposited multilayers was carried out by using Philips X'Pert diffractometer with Co-K radiation. The Knoop hardness of the deposits was measured using Future Tech Model FM-7 Hardness Tester. The microhardness was measured with a load of 25 gf with a loading time of 5 s. The hardness values reported here represent an average of five different values taken in different locations of the specimens. Residual stresses of the electrodeposited Ni/Cu multilayers were measured by XRD using grazing incidence geometry. The measurements were performed in Philips X'Pert diffractometer with Co-K line source. The details of the residual stress calculation have been mentioned in reference [21]. Tribological tests of the multilayer coatings were conducted using a ball-on-flat reciprocating wear and friction machine (Plint TE 70SLIM Micro-friction machine) using a AFBMA grade 10 tungsten carbide (WC) ball as counter body. Evaluations were carried out at five different loads of 3, 5, 7, 9 and 11 N and two different oscillating frequencies of 5 and 10 Hz. Tests were conducted for total sliding distance of 3600 mm keeping a sliding amplitude of 1 mm. Data acquisition was used to record the coefficient of friction online. After each test, the plate wear scar morphology was investigated under SEM (Philips Environmental Scanning Electron Microscope (ESEM) model XL-30 (TMP)). The ball-scar was also examined under optical microscope. The wear scar on plate was scanned using Taylor Hobson Form Telesurf Series-2 profilometer; the wear volume was calculated from the profile of the wear scar. The tests were done in air at room temperature,

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

7443

Fig. 1. XRD of multilayers electrodeposited from single solution electrolyte. tCu is 1.97 nm for all layers, and tNi are: (S1) 1.8 nm, (S2) 4.5 nm and (S3) 7.2 nm.

25 1 C with a relative humidity of about 5055%. The samples were cleaned with acetone using ultrasonic cleaner prior to the tests. The friction and wear data reported in the following sections where mean values of two replicating tests. 3. Results 3.1. X-ray Diffraction The crystal structure of the electrodeposited Ni/Cu multilayers was characterized by X-ray diffraction (XRD). Fig. 1 shows the X-ray diffraction patterns of the deposited Ni/Cu

multilayer on polished SS-plate substrate. It is clear that all the electrodeposited Ni/Cu multilayers were polycrystalline in nature and the growth of multilayers had a predominance of crystals oriented with (111) planes parallel to the substrates. And the (111) reflection happens to be the plane with highest atomic density and low surface energy planes for fcc Ni and Cu. This concludes that deposited Ni/Cu multilayer from sulphate bath on polycrystalline SS substrate had strong [111] texture. Fig. 2 shows the details of the diffraction patterns around the (111) reflection angle for S1, S2 and S3 specimens. It is to be noted that clearly discernible satellites, in all the three cases, flanked the (111) reflections indicating formation of coherent superlattice structure. The bilayer periods obtained from XRD patterns are given in Table 1. It can be seen that with the increase in tNi from 1.8 to 7.2 nm, the d-spacing of the multilayer lattice decreases. The variation of lattice parameter in multilayer structure with sublayer thickness could be similar to alloy formation obeying the Vegard's law, however, the more appropriate conclusions can be seen in the Section 3.2 from the stress free lattice parameters. The bilayer period, , obtained from EQCM experiment was found to quite well matching with the data calculated from satellite position in XRD pattern.

Fig. 2. Magnified view of (111) reflection for three samples S1, S2 and S3 showing satellite peaks (indicated by arrow mark).

Fig. 3. Lattice spacing d vs. sin2 plot for the multilayer S1 [Ni (1.8 nm)/Cu (1.97 nm)].

7444

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

3.2. Residual stress Generation of internal stress in Ni/Cu multilayer due to lattice misfit in this case is major concern because the lattice mismatch is 2.6% [lattice parameter for Cu: 3.608 , Ni: 3.517 ] [26]. Therefore, when nickel is electrodeposited on copper, it must be in tension at the Cu/Ni interface and so have a tensile stress. Conversely, if copper is electrodeposited onto nickel, it will have a compressive stress. Fig. 3 represents the d vs sin2 plot for the specimen S1, S2 and S3 varying the angle from 20 to +20 and (111) reflection had been chosen for stress measurement. The splitting of the d vs sin2 plot into two branches of opposite curvature for N 0 and b 0 indicates the presence of finite shear terms normal to the surface. The slope of the central mean line indicates the magnitude of the stress; the positive slope signifies tensile stress and negative slope depicts compressive stress within the coating. The higher the magnitude of slope the higher is the stress level within the film. It could be seen that the stress free lattice parameter, d0, obtained from the curve fitting was not far off from d value extracted from XRD peak position except for the sample S3 as mentioned in the Table 1. It might be due to the presence of low magnitude of stress in case of S1 and S2 compared to S3. However, there is a clear indication of decrease in the lattice parameter from S1 (tNi = 1.8 nm) to S2 (tNi = 4.5 nm) obeying Vegard's law. The slight exception in the order of lattice parameter from S2 to S3 could be due to anisotropic nature of the film. To elaborate more on residual stress pattern, we are including two extra data points as sample X1 [Ni (4.5 nm)/Cu(3.94 nm)] (tNi/tCu = 1.14) and X2 [Ni(4.5 nm)/Cu (7.88 nm)] (tNi/tCu = 0.57) from our previously published work [21] on the effect of Cu-sublayer thickness variation keeping tNi = 4.5 nm constant. Fig. 4 shows the variation of residual stress with Ni/Cu sublayer thickness ratio indicating the effect of both tNi and tCu. It is evident that at the beginning when nickel content was lower than copper; the stress level was

Fig. 5. The COF vs sliding distance at 5 N load and 10 Hz frequency for all the Ni/Cu multilayer films.

compressive. With the increase in tNi/tCu ratio, the residual stress changed its sign from compressive to tensile and started increasing and reached maximum tensile stress of 258.7 42.82 MPa at tNi/tCu ratio of 3.65. This particular fact is similar to deposition of Ni on Cu substrate. On the other hand, in case of tCu variation, the tensile stress was found to be started increasing very fast and the maximum tensile stress attained was 789.7 55.95 MPa at tNi/tCu ratio of 1.14. With further increase in Cu layer thickness, the tensile stress started decreasing and became high compressive at tNi/tCu ratio of 0.57 and this particular observation corroborates the origin of compressive stress. Very careful observation of stress plot (Fig. 4) shows that the maximum tensile stress is obtained at nearly equal thickness of both the components i.e. for sample with tNi = 4.5 nm and tCu = 3.94 nm. At these layer thicknesses, after initial discontinuous coverage of underneath layer, a full coverage with continuous layer structure could be obtained due to isolated grain coalescence. Therefore, the surface tension at the touched asperities [27] of the growing crystals within two adjacent layers would be at its maximum state. So a very high tensile state is reached. At higher thickness of Cu sublayer, the situation is just like copper is depositing on Ni and resulting stress becomes compressive. However, the presence of low stress in case of sample with tNi N tCu could be due to arising of interfaces and multilayer effect. 3.3. Tribological studies 3.1. Coefficient of friction (COF) Fig. 5 shows typical graphs of the variation of coefficient of friction (COF) for all the multilayer coatings with sliding distance at 5 N and 10 Hz. After establishing the initial sliding amplitude due to stiction force [28], the COF was stabilized at a steady state value. During initial phase of sliding, COF was reached maximum for S2 while it was lowest for S3. With

Fig. 4. Residual stress vs tNi/tCu plot showing the variation of stress. tNi and tCu indicates variation of stress with Ni and Cu sublayers keeping tCu = 1.97 nm and tNi = 4.5 nm constant respectively. Note: two data points are included from our previously published work, Ghosh et al [21].

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

7445

Fig. 6. The variation of average COF with load for all the electrodeposited multilayers.

the progression of sliding distance the COF of S2 reduced marginally from 0.45 to 0.42 and ultimately became equal to S1 at larger sliding distances. In case of S3, the rise of COF during the initial sliding was slightly delayed as compared to other

multilayer coatings and with the continuation of sliding and the stabilized COF was found to decrease monotonically except S1. This could be due to higher hardness of these multilayers with higher tNi unlike in S1 generate smaller size wear particles due to shallower ploughing action. These wear particles undergoes further grinding by rolling between the two surfaces, thus reducing the friction coefficient. A careful observation of the steady state COF plot shows sinusoidal variation of small amplitude ( 0.1) and wavelength with sliding distance. The modulation wavelength () of this variation is highest for S1 and lowest for S3. This variation can be related to stickslip behaviour of the tribological process and the lower indicates smaller percentage of adhesive contacts in normal adhesive abrasive wear [28]. The changes in stiction force from sample to sample is seemed to be related to obtained hardness data. It can be seen that the sample S3 with highest hardness showed shortest wavelength for variation of coefficient of friction. As the hardness decreases, the increases due to increased ductility. Average stabilized COF at 10 Hz for all the multilayer coatings was plotted with respect to increasing load is shown in Fig. 6. It is clear from this graph that all multilayer coatings exhibit reduction in COF with increase in load. The COF of multilayers S2, S3 and S4 exhibit similar trend keeping their

Fig. 7. Wear scar morphology of the S1 (a) at 3 N load and 10 Hz frequency; (b) at 11 N load and 10 Hz, the inset figure is the whole scar morphology and a part of it as indicated is shown here.

Fig. 8. Wear scar morphology of the of S3 (c) at 3 N load and 10 Hz frequency; (d) at 11 N load and 10 Hz, the inset figure is the whole scar morphology and a part of it as indicated is shown here.

7446

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

Fig. 9. (a) 3D profile showing the wear scar for multilayer S9 at 7 N and 10 Hz showing the two-valley one-hump morphology generated during wear process and (b) the corresponding 2D longitudinal scan across scar for calculating the wear volume.

relative position intact at all loads and the order is S4 b S3 b S2. The COF for S1 is highest at 3 N load and decreases sharply with increase in load to become lowest at 7 N load onwards. When compared with pure Ni and Cu coatings [21], S1 exhibit similar to pure Ni behaviour at low load and pure Cu behaviour at higher load. In this case, due to very small sublayer thickness of both the components, the multilayer effect is not prominent because of the proximity of alternate interfaces. Thus, COF follows pure component behaviour at respective loads. The friction coefficient of multilayer coatings (S2 to S4) although were not significantly different, has shown an interesting trend, i.e. with increase in tNi/tCu ratio, average COF decreases and it in conformity with our previous work [21]. 3.2. Wearscar morphology Fig. 7 (a and b) shows the wear scar morphology of multilayer specimen S1 at 3 N and 11 N respectively. The inset is whole wear scar and the magnified image of one edge is shown here. At 3 N, the wear scar morphology consist not only abrasive cutting marks along the sliding direction but also plastic deformation, unlike any other multilayer studied here. At higher load, edge build-up and material flow towards the end of the sliding stroke was observed similar to our earlier observations for specimens with tCu variation [21]. Fig. 8 (a and b) shows the wear scar morphology of specimen S3. It can be seen that at higher load abrasive wear marks became finer which lead to lower COF as seen earlier. Surface and associated subsurface plastic deformation was a prominent feature of the wear scar morphology for all scars examined at higher applied load. This observation corroborates the friction data with lower friction

coefficient at higher load. Therefore, wear scar examination reveals that at low load, i.e. in the elastic region, abrasive cutting and ploughing actions are the principal mode of surface and subsurface weight loss. And above critical load for plastic deformation, martial flow, edge build-up followed by crack formation at the stroke end are the primary causes of the wear process. 3.3. Wear rate (K) The wear volume of the scar in multilayer specimen was measured using 3D-topography (typical 3D-profile is shown in

Fig. 10. Wear rate of the electrodeposited coatings and the corresponding ball is shown here. Notice the difference in scale of wear rate for the ball and coatings.

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448

7447

Fig. 9), and the same was used to calculate wear rate using formula given elsewhere [29]. Ball wear scar radius was measured using optical microscope and was used to calculate its wear volume. Wear rate (K) of the multilayer coatings and the corresponding balls are plotted against load for all the specimens as shown in Fig. 10. It can be seen that with the increase in load, the wear behaviour of the multilayer coatings was changed. All coatings exhibit trend of increase in wear rate with load till 7 N. With further increase in load the trend of wear rate varies from coating to coating as plastic deformation plays a major role in the wear process. It is to be noted that at all loads, the trend of wear rate, KS3 KS2 b KS4 b KS1 is maintained with little variation. For S1 and S3 it increases slightly from load 7 N to 9 N and for S2 and S4, the wear rate reduces marginally for the same range. Further increase in load i.e. at 11 N, wear rate for S1 has reached to very high value as compared to other multilayers for which it has reduced marginally. It can also be seen that the wear rate for S2 and S3 were comparatively lower than S1 and S4. Wear rate of S3 is lowest at all loads while wear rate of S1 is highest at all loads. Interesting features were observed in 3-D profile pattern where wear scar at 3 N and 11 N were uniform in cross section; however, at 7 N, the profile showed deep valleys at the start and end of the stroke and shallower at the centre of the stroke. This is shown in Fig. 9. SEM investigation revealed that at low load abrasive wear was predominant while at higher load plastic flow plays major role. It seems that 7 N is the transition load; where abrasive and plastic flow both mechanisms are equally operative, resulting in two-valley one-hump 3-D profile [30]. Fig. 10 also shows the corresponding ball wear rate for all the multilayer coatings with applied load. It can be seen that ball wear rate was almost one to two order of magnitude lower than the respective coating wear rate, therefore measurement accuracy was limited. It is found that the ball wear rate increases with the increase in load and beyond 7 N a sudden increased was noticed. For S1, at low load, the wear rate was low. But the sudden increase beyond 7 N is probably because of embedment of hard worn-out particle (WC) on softer matrix lead to higher ball wear. On the other hand, at 5 N load onwards the higher ball wear rate for sample S2 was observed, which can be attributed to higher adhesion with coating. 4. Discussions In this section, the observed wear rate data is explained in terms of (H/E) ratio, i.e. in terms to elastic strain to failure as discussed previously by Leyland and Matthews [19,20]. The order of hardness as obtained from Table 1 is HS3 N HS2 N HS4 N HS1. The elastic modulus data for these specimens can be approximated from Tench and White's tensile data [15] for multilayers, E values of pure Ni (200 GPa) and pure Cu (124 GPa). The approximated elastic modulus order is ES4 N ES3 N ES2 N ES1. So, the order of elastic strain to failure is (H/E)S3 N (H/E)S2 N (H/E)S4 N (H/E)S1. In fact, the calculated wear rate as seen in Section 3.3.3 follows the identical order. At 35 N load, the wear rate order is KS3 KS2 b KS4 b KS1,

and at 711 N load, the order is KS3 b KS2 b KS4 b KS1. Therefore, it is obvious that the elastic strain to failure (H/E) ratio is equally applicable in explaining the wear behaviour of the metallic multilayer coatings along with the hardness. Moreover, in case of S1 with tNi: tCu 2:2, the multilayer effect is not prominent due to diffuse interface pattern and repulsing effect of the two close by interfaces; therefore, strengthening due to interface effect (Koehler effect) is poor and shows more like an alloy type of behaviour. This effect is also reflected in its hardness value. With progressive increase in thickness of Ni layer i.e. at 5 and 8 nm, the strengthening due to multilayer effect was significantly increased and thus attributing lower wear rate. This mechanism has resulted in the increase hardness of these samples. With further increasing Ni layer thickness, in case of S4 with 11 nm Ni, the strengthening due to dislocation barrier at interface is getting reduced because of allowance of more room for dislocation motion within a relatively thick Ni layer and thus gets reflected in slightly lower hardness value. Even though the stress level in S1 is compressive in nature ( 107.7 19.25 MPa), but its low elastic strain to failure increases its ductility. Therefore, even at low load, it undergoes plastic deformation with increased wear process. However, in case of specimen S3 with highest (H/E) ratio, the presence of small tensile stress (258.7 42.82 MPa) is found to be not much detrimental to wear loss at medium and higher loads. At 3 N load, when the wear process is exclusively abrasive in nature, the higher tensile stress for S3 compared to S2 (110.5 36.34 MPa) led to subsurface spalling and delamination during microscopic abrasive wear process. That is why at low load, the wear rate of S3 becomes almost equal to S2, i.e. KS3 KS2. It indicates that higher tensile stress also might have contributed to the higher wear volume. The present study highlights the fact that previous investigators [9,10,22,23] concentrated on wear studies of Ni/Cu multilayers mainly with tNi = tCu without knowing their implication on internal stress and subsequently detrimental effect on wear rate. It also emphasizes the fact that the wear resistance of Ni/Cu multilayers with unequal thickness of component layers is better owing to presence of low tensile stress. Therefore, this study confirms that minimum wear rate can be obtained at intermediate tNi (58 nm) keeping constant tCu at 2 nm. 5. Conclusions Ni/Cu multilayers with varied tNi, keeping tCu 2 nm constant, have been electrodeposited from sulphate based electrolyte. XRD analysis confirmed the formation of homogeneous and compact multilayer structure. The measured residual stress was found to be dependent on tNi/tCu ratio and reached its highest magnitude (tensile) at nearly equal values of tCu and tNi which was found to be harmful to wear process. The detailed analysis of COF results indicated that multilayer sample with tNi tCu 2 nm showed COF behaviour of Ni at low loads and that of Cu at higher loads. The COF behaviour for other multilayers maintained the order S4 [Ni(11 nm)/Cu(1.97 nm)] b S3 [Ni(7.2 nm)/Cu(1.97 nm)] b S2 [Ni(4.5 nm)/Cu(1.97 nm)] at all loading conditions and showed that with increase in tNi/tCu

7448

S.K. Ghosh et al. / Surface & Coatings Technology 201 (2007) 74417448 [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] E.O. Hall, Proc. Phys. Soc. B 64 (1951) 747. N.J. Petch, J. Iron Steel Inst. 173 (1953) 25. W.D. Nix, Scr. Mater. 39 (1998) 545. D.M. Tench, J.T. White, J. Electrochem. Soc. 138 (1991) 3757. D.M. Tench, J.T. White, Met. Trans. A 15A (1984) 2039. H.C. Barshilia, K.S. Rajam, Surf. Coat. Technol. 155 (2002) 195. A. Mishra, M. Verdier, Y.C. Lu, H. Kung, T.E. Mitchell, N. Natasi, J.D. Embury, Scr. Mater. 39 (1998) 555. A. Leyland, A. Matthews, Wear 246 (2000) 1. A. Leyland, A. Matthews, Surf. Coat. Technol. 177178 (2004) 317. S.K. Ghosh, P.K. Limaye, B.P. Swain, N.L. Soni, R.G. Agrawal, R.O. Dusane, A.K. Grover, Surf. Coat. Technol. 201 (2007) 4609. W. Zhang, Q. Xeu, Thin Solid Films 305 (1997) 292. W. Zhang, Q. Xeu, Wear 214 (1998) 7. A.W. Ruff, N.K. Myskin, Trans. ASME, J. Tribol. 111 (1989) 156. A.S.M.A. Haseeb, J.P. Celis, J.R. Roos, Thin Solid Films 444 (2003) 199. T.E. Such, Trans. Inst. Met. Finish. 1 (4) (1992) 21. S.K. Ghosh, A.K. Grover, M.K. Totlani, Trans. Inst. Met. Finish. 80 (2) (2002) 56. F.E. Kennedy, C.A. Brown, J. Kolodny, B.M. Sheldon, Trans. ASME, J. Tribol 121 (4) (1999) 968. K. Holmbug, A. Matthews, Coating Tribology, Elsevier, Amsterdam, 1994. S. Fouvry, P. Kapsa, H. Zahouani, L. Vincent, Wear 203204 (1997) 393.

ratio the average COF decreased. Coating wear rate was seen to increase with load generally and stabilised beyond 7 N. Ball wear rate also increased with load for all coatings. The wear rate of the multilayer was minimum for intermediate tNi ( 5 8 nm). The magnitude of wear rate for coatings, with lower and higher tNi, was high due to diminishing multilayer effect. The order of wear rates for multilayers was found to follow as the order of their (H/E) ratio and hardness. References
[1] L.W. Wang, Y. Liu, Z. Zhang, Handbook of Nanophase and Nanostructured Materials, Kluwers Academic Publishers, Dordrecht, 2002. [2] N. Sulitanu, Mater. Sci. Eng. B 95 (2002) 230. [3] T. Osaka, Electrochim. Acta 45 (2000) 3311. [4] S.L. Lehozky, Phys. Rev. Lett. 41 (1978) 1814. [5] J.F. Archard, J. Appl. Phys. 24 (1953) 981. [6] S. Veprek, Thin Solid Films 317 (1998) 449. [7] J. Musil, Surf. Coat. Technol. 125 (2000) 322. [8] S. Veprek, Thin Solid Films 476 (2005) 1. [9] A.W. Ruff, Z. Wang, Wear 131 (1989) 259. [10] A.W. Ruff, D.S. Lashmore, Wear 151 (1991) 245. [11] J.S. Koehler, Phys. Rev. B 2 (1970) 547.

Das könnte Ihnen auch gefallen