You are on page 1of 7

Applied Surface Science 254 (2008) 59105916

Contents lists available at ScienceDirect

Applied Surface Science


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

Effect of pulse electrodeposition parameters on the properties of Ni/nano-SiC composites


P. Gyftou, E.A. Pavlatou, N. Spyrellis *
Laboratory of General Chemistry, School of Chemical Engineering, National Technical University of Athens, 9, Heroon Polytechniou Street, Zografos Campus, Athens GR-15780, Greece

A R T I C L E I N F O

A B S T R A C T

Article history: Received 15 January 2008 Received in revised form 21 March 2008 Accepted 23 March 2008 Available online 1 April 2008 PACS: 81.15.Pq 81.05.Ni 82.45.Yz 68.55.Jk 62.20.Qp Keywords: Nickel electrodeposition Composite Silicon carbide Pulse plating Preferred orientation Microhardness

Pure nickel and nickel matrix composite deposits containing nano-SiC particles were produced under both direct and pulse current conditions from an additive-free nickel Watts type bath. It has been proved that composite electrodeposits prepared under pulse plating conditions exhibited higher incorporation percentages than those obtained under direct plating conditions, especially at low duty cycles. The study of the textural perfection of the deposits revealed that the presence of nano-particles led to the worsening of the quality of the observed [1 0 0] preferred orientation. Composites with high concentration of embedded particles exhibited a mixed crystal orientation through [1 0 0] and [2 1 1] axes. The embedding SiC nano-particles in the metallic matrix by an intra-crystalline mechanism resulted in the production of composite deposits with smaller crystallite sizes and more structural defects than those of pure Ni deposits. A dispersion-hardening effect was revealed for composite coatings independently from applied current conditions. Pulse electrodeposition signicantly improved the hardness of the Ni/SiC composite deposits, mainly at low duty cycle and frequency of imposed current pulses. 2008 Elsevier B.V. All rights reserved.

1. Introduction Electroplating is a method of codepositing micron- or nanosized particles of metallic or non-metallic compounds and polymers with a metal or alloy matrix. Composite deposits are used in various elds, from high-tech industries such as electronic components and computers, to more traditional industries such as general mechanics and automobiles, paper mills, textiles and food industries. During the last decades, the main work carried out in this eld is aimed almost entirely to the production of wear- and corrosion-resistant coatings, self-lubricating systems and dispersion-strengthened coatings [13]. For nickel matrix electrodeposits in particular, a great variety of particles have been used such as hard oxides (like SiO2, Al2O3, TiO2), carbides like WC [4] and SiC [5 8], liquid containing polymeric microcapsules [9,10], carbon nanotubes [11], etc. With the increasing availability of nano-particles, the interest of the low-cost and low-temperature composite

* Corresponding author. Tel.: +30210 7723085; fax: +30210 7723088. E-mail address: nspyr@chemeng.ntua.gr (N. Spyrellis). 0169-4332/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2008.03.151

electroplating is continuously growing, with major challenge being the achievement of high codeposition rates and homogenous distribution of the particles in the metallic matrix. A recent review on the electrodeposition of metal matrix composite coatings containing nano-particles can be found in literature [12]. Silicon carbide nickel matrix coatings have been extensively studied and have gained widespread use for the protection of friction inside parts of cylinders in the automotive industry. Considerable research has been mainly focused on the impact of the electrodeposition parameters: electrolysis conditions (composition of the electrolytic bath, presence of additives, pH value), current conditions (type of imposed current and values of the current density) and properties of the reinforcing particles (size, surface properties, concentration and type of dispersion in the bath) on the electrolytic codeposition process of SiC with Ni, as well as the properties of the composite coatings [58,1222]. In general, it has been observed that the amount of embedded SiC particles increases with both increasing concentration of suspended SiC particles and of additives presence in the electrolyte [5,19,20]. It has also been reported that a decrease in the SiC particle size affects the wear and corrosion resistance in a positive

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916

5911

way. These phenomena are mainly attributed to the hardening of the metal matrix by nely dispersed particles, which weaken the plowing effect and adhesive wear and retard the grain growth in the matrix at elevated temperatures, so good mechanical properties can be sustained [13,16,20]. However, the particles should be smaller than roughly a micron and must be uniformly dispersed to exhibit the dispersion-hardening effect [23]. But the reduction of particle size decreases the codeposition content of the particles [5,7,14,24]. Thus, it seems that the particle size plays an important role in the electro-codeposition process and hence in the mechanical properties of the Ni/SiC composite coatings. Moreover, recent research has pointed out that the physico-chemical properties of SiC particles are crucial to the understanding of the codeposition mechanism [7,8,17,21,24]. Relatively less research work has been done to examine the effect of the current type on the codeposition process of the composite Ni/SiC electrocoatings [7,8,1214,18,22], although a signicant number of reports have shown that the microstructure and properties of pure Ni deposits may be effectively altered under specic pulse current conditions [25,26]. Pulse electrodeposition has been found to be an effective means of perturbing the adsorptiondesorption phenomena occurring at the electrode electrolyte interface and hence the electrocrystallization process. In addition, pulse plating permits higher current densities than the limiting direct current density to be attained and thus nanocrystalline Ni deposits could be produced [17,25]. In the case of Ni/ SiC codeposits, the application of pulse current techniques results in the production of composite coatings with higher percentages of incorporation, reduced grain sizes and a more uniform distribution of SiC particles in the Ni matrix than those attained under direct current regime [68,12,13,18]. It is noteworthy that, as mentioned above, although there is a plenty of research work concerning either the effect of operating conditions on the amount of codeposited SiC particles or/and the resulting mechanical properties of composite coatings, the effect of the embedded particles on the microstructure of the coatings and, consequently, the relation between structure and properties has been investigated to a limited extent [7,8,16,21]. In addition, it is only recently that a few papers have reported the incorporation of nano-sized particles (and SiC in particular) into metal matrices [5,19,22,27]. In the present study, the electrolytic codeposition of nano-sized SiC particles in Ni matrix from an additive-free nickel Watts solution by applying both direct and pulse electroplating was investigated. The codeposition percentage of SiC, the preferred orientation of nickel crystallites, the microstructure and the microhardness of the coatings were determined. Pure Ni deposits were also produced under the same experimental conditions for comparison. 2. Experimental Pure Ni and composite Ni/SiC coatings were electrolytically deposited from an additive-free nickel Watts type bath. The electrodeposition experiments were performed on rotating disc electrodes (RDEs) with a rotation velocity of 200 rpm, under either direct or pulse current conditions. In the case of Ni/SiC composite deposits a commercial SiC powder with a mean diameter of 20 nm (b-SiC, Marketech International INC) was added into the bath. The SiC particles were maintained in suspension by continuous agitation by the use of a 250 rpm magnetic stirrer placed at the cell bottom, for at least 12 h before codeposition. The substrates were brass discs with a mirror nish obtained by mechanical polishing and were chemically cleaned in an ultrasonic agitated bath before deposition. The anode, a nickel foil of 99.9% purity, was positioned on the side of the electrolytic cell.

Table 1 Overview of the electrodeposition parameters for preparation of pure Ni and composite Ni/SiC coatings Solution composition Electrolyte (Watts type) NiSO47H2O NiCl26H2O H3BO3 SiC powder (mean diameter of 20 nm) Electrodeposition conditions pH Temperature Substrate Cathode rotation rate (v) Electrolyte agitation Anode Current density Current type Duty cycle (d.c.) Pulse frequency (v)

300 g L1 35 g L1 40 g L1 20 g L1 4.40 50 1 8C Brass disc (diameter 25 mm) 200 rpm Magnetic stirring (250 rpm) Ni foil 5 A dm2 Direct (DC) or pulse (PC) 10, 50 and 90% 1, 10 and 100 Hz

During electrodeposition the temperature of the plating solution was maintained at 50 1 8C and the initial pH of the electrolyte was adjusted to a constant value of 4.4. In pulse deposition experiments the frequency (v) of the imposed pulses varied between 1 and 100 Hz, while the duty cycle (d.c.) of the pulses varied between 10 and 90%. It should be noticed that d.c. = Ton/(Ton + Toff), where Ton is the time period when the pulses are imposed and Toff is the relaxation time. The composition of the plating solution with a 1 L total volume, as well as the deposition parameters, is given in Table 1. Ultrasonic cleaning in de-ionized water for 10 min of the deposits followed the electrolytic codeposition so as to remove any loosely adsorbed SiC particles from their surface. The thickness of the produced coatings was at least 40 mm in order to obtain preferred orientations fully developed and therefore permit their determination in such a way to attain results comparable and reliable. Measurements of the Vickers microhardness (HV in kgf mm2) of pure nickel and Ni/SiC composite deposits were performed on their surface by using a Reichert microhardness tester under 50 g load for a period of 10 s and the corresponding nal values were determined as the average of 10 measurements. The thickness of the coatings and the selected load were chosen so as to avoid any substrate effect on the microhardness value measured [28]. The preferred crystalline orientation of Ni was examined by applying X-ray diffraction (XRD) technique utilizing a Siemens D5000 diffractometer with a Cu-Ka radiation. The characterization of the deposits morphology and microstructure was performed by scanning electron microscopy (SEM, JSM-6300) and transmission electron microscopy (TEM, Philips CM12), while the amount of embedded SiC particles was evaluated by the use of the energy dispersive X-ray spectroscopy (EDS, Oxford Instruments). The determined amount of codeposited SiC particles in the Ni matrix resulted from the mean value of at least three measurements for each deposit. 3. Results and discussion 3.1. SiC particles codeposition The incorporation percentages (vol.%) of nano-SiC particles in Ni matrix composites, prepared under both direct (DC) and pulse plating (PC) conditions, as a function of duty cycle at the selected frequencies, are shown in Fig. 1. From this gure it was concluded that composite electrodeposits prepared under PC conditions had higher incorporation percentages than those obtained under DC conditions. The highest percentage was observed at d.c. = 10% and v 10 Hz.

5912

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916

Fig. 1. Volumetric percentage and number density of SiC particles codeposited with Ni at different pulse frequencies vs. duty cycle (d.c.). Data for DC conditions are denoted as a bar for d.c. = 100%.

Examining the experimental results, it was concluded that the highest codeposition percentage was achieved at the lowest duty cycle, i.e. d.c. = 10% for each frequency. This remark could be linked to the suggestion of a previous research that the maximum particle

concentration is achieved where the deposits thickness per cycle approaches the particles diameter size [29]. In the present study, as the mean diameter of the SiC particles used is 20 nm, the deposition time could be short and the crucial factor is the rest period that permits the enrichment of the catholytic area with particles. According to previous results of ours, this tendency was also observed in the case of micron-sized SiC particles codeposition under similar PC experimental conditions [8]. In this point, it should be noticed that several researchers who studied only micron-sized particles had reported that the reduction of particle size decreased the codeposition percentages [15] and the smaller the particle size was, the more difcult the particles were embedded in the metallic matrix [30]. However, regarding the determination of the incorporation percentage of the nano-SiC particles and its comparison with micron-sized ones, especially when there is a need to study the effect of particles size on various composite deposits properties, the gravimetric or volumetric percentage is not sufcient and the number density of codeposited particles must be taken into account [6,8,15,3132]. Converting the volumetric percentage of nano-SiC particles to the corresponding number density in the composite coatings, assuming that the particles are mono-dispersed, spherical and uniformly distributed in the coatings, the codeposition rate varied between 7.5 1015 and 18.5 1015 particles cm3 for PC conditions, while for DC conditions the codeposition rate was 7.3 1015 particles cm3, as depicted in Fig. 1. The pronounced increase of the number of the embedded particles under PC conditions was in several cases three times more than under DC conditions, and this observation could explain the inuence of the codeposited SiC nano-particles on the structure and the mechanical properties of the deposits.

Fig. 2. Quality of [1 0 0] crystalline orientation for: (a) pure Ni and (b) composite Ni/ SiC deposits, prepared at different pulse frequencies vs. duty cycle (d.c.). Data for DC conditions are denoted as a bar for d.c. = 100%.

Fig. 3. XRD patterns of: (a) a pure Ni and (b) a Ni/SiC coating, both prepared under the same PC conditions (d.c. = 90% and v 100 Hz).

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916

5913

3.2. Crystalline orientation For the determination of the deposits preferred crystalline orientation, as well as for the evaluation of the quality of this orientation the term Relative Texture Coefcient (RTC(hkl)) was used, dened as: Ishkl =Iphkl 100% RTChkl P6 i1 Ishkl =Iphkl where Is(hkl) and Ip(hkl) are the diffraction intensities of the (hkl) plane measured in the diffractogram for the deposit and the standard Ni powder sample, respectively. Only the six basic of the total 8 reection lines for Ni have been considered, i.e. (1 1 1), (2 0 0), (2 2 0), (3 1 1), (3 3 1) and (4 2 0), since the diffraction lines of (2 2 2) and (4 0 0) are the second-order diffraction of (1 1 1) and (2 0 0) planes, respectively. It is of interest to note that, when using a Cu-Ka radiation, the reinforcement of the lines (3 1 1) and (1 1 1) is attributed to a dispersed [2 1 1] orientation [33]. The term RTC(hkl) expresses the percentage of the relative intensity of a given orientation (hkl) among the 6 crystallographic orientations of each Ni and Ni/SiC sample studied, while a preferred orientation of hkl plane is indicated by values of RTC(hkl) ! 16.67% [34]. The XRD diagrams of all pure Ni and composite Ni/SiC deposits prepared under both DC and PC conditions revealed a [1 0 0] preferred orientation but, depending on the current parameters,

the quality of [1 0 0] orientation varied. The results for the evaluation of the quality of [1 0 0] orientation for pure Ni and composite Ni/SiC deposits are presented in Fig. 2a and b, respectively. It appeared that, especially for composites, the application of PC results to the worsening of the [1 0 0] textural perfection compared to that under DC conditions. Moreover, the SiC nano-particles embedding leads in general to the worsening of the [1 0 0] texture in comparison to the pure Ni electrodeposits prepared under similar PC experimental conditions, as depicted in Figs. 2a and b and 3a and b. This textural worsening was accompanied by the reinforcement of [2 1 1] crystalline orientation, as illustrated in Fig. 3a and b. In this point, it should be noticed that the reinforcement of [2 1 1] crystalline orientation is concluded by the relative increasing of (3 1 1) and (1 1 1) intensities [33]. This tendency was also observed in the case of Ni/micron-SiC deposits studied in our previous work [8]. Hence, it was revealed that the embedding of the SiC particles in the Ni matrix presented a tendency to modify the [1 0 0] texture to a mixed orientation of nickel crystallites through [1 0 0] and [2 1 1] axis. This effect could be attributed not only to the engulng of the SiC particles in the Ni matrix, but also to their physico-chemical interaction when approaching the catholytic area. The SiC particles intervene by changing the catholyte composition by adsorption desorption phenomena of H+ on their surface [8,35], resulting to an increase of the [2 1 1] mode of crystal growth, which is attributed to the presence of colloidal dispersion of nickel hydroxide in the case of pure nickel electrodeposition [33,35,36]. Consequently, for the composite electrodeposits the quality of [1 0 0] orientation

Fig. 4. SEM images of (a) a pure Ni deposit and (b) a Ni/SiC deposit, both prepared under the same PC conditions (d.c. = 50% and v 10 Hz).

Fig. 5. (a) SEM image of the surface and (b) the corresponding Si mapping image of a Ni/SiC deposit prepared under PC conditions (d.c. = 10% and v 1 Hz). The volumetric percentage of embedded SiC particles is 6.3%.

5914

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916

depends on the quantity of the embedded particles and the applied PC conditions. 3.3. Surface morphologySiC particles dispersion The microstructure of conventional electrodeposits consists of columnar grains, whose diameters increase with increasing the deposits thickness [37]. Such deposits show a very strong crystallographic texture due to the preferential growth of favorably oriented crystals and the size of these grains is usually above the nano-regime. Therefore, regarding the present study, the highest value of the texture perfection is where the pure nickel deposits consisted by very well formed, micron-sized [1 0 0] crystallites [3]. In order to form smaller or even nano-sized crystals, continuous nucleation of new crystals should occur and this can be achieved by: (a) adding suitable grain rening agents [38] to the electrodeposition bath, (b) adding a reinforcing second phase [39], and/or (c) pulsating the deposition current [36]. Previous studies have shown that the embedding of SiC particles in the Ni matrix changes the shape of Ni crystallites from columnar to equiaxial and reduces their average grain size, while their orientations approach a rather random distribution [5,37]. On the contrary, according to our results, randomly oriented composite deposits were not observed and the lowest value of texture perfection should be related to the weakly formed, very small sized [1 0 0] crystallites and the formation of [2 1 1] oriented microcrystallites [38].

Having these remarks in mind, it was expected that the embedding of SiC nano-particles resulted in the production of composite deposits with smaller crystallite sizes than those of pure Ni deposits prepared under similar conditions. The reduction of Ni crystallites grain size due to the presence of SiC particles is discernible in the SEM micrographs of composite deposits, as shown in Fig. 4b, compared with those of pure Ni deposits prepared under similar conditions as depicted in Fig. 4a. The comparison of these images demonstrated that the surface of pure Ni deposit consisted by very well formed crystallites with grain size that exceeded few microns (Fig. 4a). On the other hand, composites exhibited a cauliower structure that was formed by microcrystalline aggregates (Fig. 4b). In addition, it seemed that increased SiC codeposition was linked to more microcrystalline structures, as illustrated in Figs. 5a and 6a. This microcrystalline structure was also veried by TEM micrographs (Fig. 7), where some large typical [1 0 0] crystals characterized by oblique parallel twin planes surrounded by a microcrystalline magma [40] are apparent. Moreover, the dispersion of nano-SiC particles on the surface was rather uniform, and appeared that high frequency values of pulse current favored the dispersion of SiC particles as presented in Figs. 5b and 6b. Recent investigations have shown that, when codepositing micron-sized SiC particles in a Ni matrix, large columnar Ni crystals grow undisturbed between the particles, or as far as they are not terminated by a particle, and the layer growth starts with a new nano-sized initial layer only on the top of the particles [39]. However, in the case of nano-sized SiC particles, the phenomenon is quite different as the SiC particles due to their very small size could be incorporated not only at the borders and the edges of the nickel crystallites, but also inside the nickel crystals, leading to an intra-crystalline codeposition mechanism [8]. Therefore, the presence of SiC nano-particles inside the Ni crystals resulted to enhanced structural defects.

Fig. 6. (a) SEM image of the surface and (b) the corresponding Si mapping image of a Ni/SiC deposit prepared under PC conditions (d.c. = 10% and v 100 Hz). The volumetric percentage of embedded SiC particles is 7.2%.

Fig. 7. TEM image of a Ni/SiC deposit prepared under PC conditions (d.c. = 10% and v 100 Hz).

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916

5915

Composite Ni/SiC deposits prepared under either direct or pulse current conditions exhibited higher HV values than pure Ni deposits produced under similar plating conditions (Fig. 8). It is known that the hardness and other mechanical properties of metal matrix composites the phrase (MMCs) depend not only on the mechanical characteristics of the matrix, particles and interfaces, but also on the amount and size of the dispersed phase. The amount and size of particles dene two kinds of reinforcing mechanisms in MMCs, namely dispersion-strengthening and particle-strengthening [15]. In our case, a dispersion-strengthening effect was achieved, therefore increased hardness values are obtained at lower volumetric percentage of codeposited nano-SiC particles than in the case of micron-SiC particles [15,32,41]. Fig. 8b demonstrates a signicant dependence of the HV values of composite deposits on the duty cycle and pulse current frequency imposed. Enhanced hardness was achieved at low duty cycle and low pulse frequency. The observed hardening effect of the pulse current regime could be related to the formation of ne nickel grains (Fig. 7) obtained at pulse-off times longer than pulseon times [8,27,28]. Additionally, the intra-crystalline embedding mechanism of the nano-SiC particles and, consequently, the structural modication of Ni crystallites accompanied by increased number of crystalline defects expressed through the worsening of the preferred [1 0 0] textural perfection could be associated with the observed high HV values. 4. Conclusions Ni matrix composite electrodeposits containing nano-SiC particles were produced under both direct and pulse current conditions from an additive-free nickel Watts type bath. It has been proven that composite electrodeposits prepared under PC conditions had higher incorporation percentages than those obtained under DC conditions: generally the percentage of embedded SiC particles increased with decreasing duty cycle. The application of PC plating led to the worsening of the [1 0 0] textural perfection in comparison with DC plating for both pure and composite coatings. The embedding of SiC nano-particles resulted also to an intensive worsening of the [1 0 0] textural perfection compared to pure Ni deposits prepared under similar PC conditions, in such an extent that a mixed crystal orientation through [1 0 0] and [2 1 1] axis was imposed. The presence of SiC nano-particles in the metallic matrix resulted in the production of composite deposits with smaller crystallite sizes and more structural defects than those of pure Ni deposits. In addition, the nano-particles embedded by an intracrystalline mechanism demonstrated a rather uniform dispersion on the surface. A dispersion-hardening effect was revealed for composite Ni/ SiC deposits independently from applied current conditions compared to Ni deposits. The application of PC regime increased the microhardness of composite deposits and enhanced HV values were obtained at low duty cycles and low frequencies. The observed hardening effect under PC deposition could be attributed to the formation of ne grains of Ni matrix, the embedding mechanism of nano-SiC particles, as well as the structural modications induced by the presence of the particles and the applied current conditions. References
[1] J. Fransaer, J.P. Celis, J.R. Roos, Met. Finish. 91 (1993) 97. [2] M. Musiani, Electrochim. Acta 45 (2000) 3397. [3] M. Srivastava, V.K.W. Grips, A. Jain, K.S. Rajam, Surf. Coat. Technol. 202 (2) (2007) 310.

Fig. 8. Vickers microhardness values of: (a) pure Ni and (b) composite Ni/SiC deposits prepared under DC and PC conditions at different pulse frequencies vs. duty cycle. Data for DC conditions are denoted as a bar for d.c. = 100%.

Taking all the above mentioned into account, the pulsed plated composite deposits were characterized by smaller nickel crystallites with more structural defects in comparison to pure ones, which could be related to the observed worsening of their RTC values. Consequently, higher hardness values were expected for the pulse plated composites. 3.4. Hardness Vickers microhardness (HV) values of the pure Ni and composite Ni/SiC deposits, prepared under both DC and PC conditions, are shown in Fig. 8a and b, respectively. The hardness values of pure Ni deposits varied between $215 and $265 kgf mm2, while the corresponding values for Ni/SiC deposits varied between $275 and $850 kgf mm2. The hardness of pure Ni deposits prepared in the PC regime was always higher than the corresponding value of the deposit prepared under DC conditions and attained the highest values at low duty cycles for every pulse frequency. At such prolonged relaxation times, a reinforced adsorption of anions (or other species) from the catholyte to the metallic surface has been proposed [25]. These adsorption phenomena provoked re-nucleation processes during the next on-pulses and consequently grain renement and enhanced structural defects were imposed, and thus leading to higher HV values of the deposits.

5916

P. Gyftou et al. / Applied Surface Science 254 (2008) 59105916 [22] F. Hu, K.C. Chan, Appl. Surf. Sci. 233 (2004) 163. [23] P.G. Shewmon, Transformations in Metals, McGraw-Hill, New York, 1969, pp. 317319. [24] H.-K. Lee, H.-Y. Lee, J.-M. Jeon, Surf. Coat. Technol. 201 (8) (2007) 4711. [25] C. Kollia, N. Spyrellis, J. Amblard, M. Froment, G. Maurin, J. Appl. Electrochem. 20 (1990) 1025. [26] A.M. El-Sherik, U. Erb, J. Page, Surf. Coat. Technol. 88 (1996) 70. [27] M. Sarret, C. Muller, A. Amell, Surf. Coat. Technol. 201 (1-2) (2006) 6352. [28] H.E. Boyer (Ed.), Hardness Testing, ASM Intrenational, Metals Park, OH, 1987. [29] E.J. Podlaha, D. Landolt, J. Electrochem. Soc. 144 (7) (1997) L200. [30] H.K. Lee, H.Y. Lee, J.M. Jeon, Surf. Coat. Technol. 201 (8) (2007) 4711. [31] O. Berkh, A. Bodnevas, J. Zahavi, Plat Surf. Finish. 11 (1995) 62. [32] C.F. Malfatti, J. Zoppas Ferreira, C.B. Santos, B.V. Souza, E.P. Fallavena, S. Vaillant, J.P. Bonino, Corros. Sci. 47 (2005) 567. [33] J. Amblard, M. Froment, N. Spyrellis, Surf. Technol. 5 (1977) 205. [34] X. Xe, J.P. Celis, M. De Bonte, J.R. Roos, J. Electrochem. Soc. 141 (10) (1994) 2698. [35] C.S. Lin, K.C. Huang, J. Appl. Electrochem. 34 (2004) 1013. [36] J. Amblard, I. Epelboin, M. Froment, G. Maurin, J. Appl. Electrochem. 9 (1979) 233. [37] F. Ebrahimi, G.R. Bourne, M.S. Kelly, T.E. Mathews, Nanostruct. Mater. J. 11 (3) (1999) 343. [38] E.A. Pavlatou, M. Raptakis, N. Spyrellis, Surf. Coat. Technol. 201 (2007) 4571. [39] T. Lampke, B. Wielage, D. Dietrich, A. Leopold, Appl. Surf. Sci. 253 (5) (2006) 2399. [40] N. Spyrellis, J. Amblard, M. Froment, G. Maurin, J. Microsc. Spectrosc. Electron. 12 (1987) 221. [41] R. Mishra, B. Basu, R. Balasubramaniam, Mater. Sci. Eng. A 373 (2004) 370.

[4] M. Stroumbouli, P. Gyftou, E.A. Pavlatou, N. Spyrellis, Surf. Coat. Technol. 195 (2005) 325. [5] S.-C. Wang, W.-C.J. Wei, Mater. Chem. Phys. 78 (2003) 574. [6] P. Gyftou, M. Stroumbouli, E.A. Pavlatou, P. Asimidis, N. Spyrellis, Electrochim. Acta 50 (2005) 4544. [7] P. Gyftou, M. Stroumbouli, E.A. Pavlatou, N. Spyrellis, Trans. Inst. Met. Finish. 80 (3) (2002) 88. [8] E.A. Pavlatou, M. Stroumbouli, P. Gyftou, N. Spyrellis, J. Appl. Electrochem. 36 (2006) 385. [9] S. Alexandridou, C. Kiparissides, J. Fransaer, J.P. Celis, Surf. Coat. Technol. 71 (1995) 267. [10] A. Kentepozidou, C. Kiparissides, F. Kotzia, C. Kollia, N. Spyrellis, J. Mater. Sci. 31 (1996) 1175. [11] X.H. Chen, F.Q. Cheng, S.L. Li, L.P. Zhou, D.Y. Li, Surf. Coat. Technol. 155 (2002) 274. [12] C.T.J. Low, R.G.A. Wills, F.C. Walsh, Surf. Coat. Technol. 201 (1-2) (2006) 371. [13] I. Garcia, J. Fransaer, J.P. Celis, Surf. Coat. Tecnhol. 148 (2001) 171. [14] L. Orlovskaja, N. Periene, M. Kurtinaitiene, S. Surviliene, Surf. Coat. Technol. 111 (1999) 234. [15] A.F. Zimmerman, G. Palumbo, K.T. Aust, U. Erb, Mater. Sci. Eng. A 328 (2002) 137. [16] I. Garcia, A. Conde, G. Langelaan, J. Fransaer, J.P. Celis, Corros. Sci. 45 (2003) 1173. [17] S.H. Yeh, C.C. Wan, J. Appl. Electrochem. 24 (1994) 993. [18] A.F. Zimmerman, D.G. Clark, K.T. Aust, U. Erb, Mater. Lett. 52 (2002) 85. [19] M.-D. Ger, Mater. Chem. Phys. 87 (2004) 67. [20] K.H. Hou, M.-D. Ger, L.M. Wang, S.T. Ke, Wear 253 (2002) 994. [21] R.P. Socha, P. Nowak, K. Laajalehto, J. Vayrynen, Coll. Surf. A 235 (2004) 45.

You might also like