Sunteți pe pagina 1din 6

Diamond & Related Materials 15 (2006) 1570 1575 www.elsevier.

com/locate/diamond

Effect of particle size on the co-deposition of diamond with nickel in presence of a redox-active surfactant and mechanical property of the coatings
Nabeen K. Shrestha, Takashi Takebe, Tetsuo Saji
Department of Chemistry and Materials Science, Tokyo Institute of Technology, Tokyo 152-0085, Japan Received 14 March 2005; received in revised form 9 November 2005; accepted 20 December 2005 Available online 14 February 2006

Abstract Particle reinforced Ni/diamond composite coatings containing various amount of the reinforcement (up to 57.6 vol.%) were prepared by the electroplating technique using a Watt's nickel bath in which the diamond particles were dispersed with the aid of a redox-active surfactant containing an azobenzene moiety. Five different sizes of diamond particles with the mean diameter ranging from 0.21 to 2.91 m were used to investigate the effect of the particle size in their co-deposition with nickel under the influence of the above redox-active surfactant and the mechanical property especially the micro-hardness and the abrasion wear of the composite coatings reinforced with various amount of the diamond particles was studied. Among all, the smallest diamond particles exhibited the highest particle co-deposition of 57.6 vol.%. The microhardness and the wear resistance of this composite with the 57.6 vol.% diamond were about 4.5 and 14 times, respectively, than that of a nickel coating without any reinforcement. 2006 Elsevier B.V. All rights reserved.
Keywords: Diamond; Nickel; Azobenzene; Surfactant; Composite plating

1. Introduction Reinforcement of various inert particles from an electrolytic bath into the metallic deposit is a well-established technique [1 3]. This technique produces a composite coating with specific mechanical and chemical properties, which cannot be obtained, with the pure plated metallic coatings without the reinforced particles. In order to enhance the mechanical and chemical properties of a composite coating so as to get its best performance, researchers are trying their best to increase the particle content in most case of the particle dispersed metal matrix composite coatings. For the advanced coating, various hard ceramic particles can be dispersed into a metal matrix by this method of composite plating. Such coatings can be used as an advanced engineering material in various applications where high anti-wear performance of the material is needed. Wear is a rather complex process, which is generally a result of mechanical, thermal and chemical interactions of the surface

Corresponding author. Tel./fax: +81 3 5734 2627. E-mail address: tsaji@o.cc.titech.ac.jp (T. Saji). 0925-9635/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.diamond.2005.12.040

with the surroundings. Due to this complexity, it is hardly possible to predict the wear resistance of a material simply based on the property of the material. However, the performance of a coating is believed to depend largely on the mechanical and chemical properties of the reinforcement. Hardness is the one of the mechanical property, which is mostly used to characterize the wear resistance of a material, and a hard coating is supposed to have a good anti-wear performance. Diamond is the hardest martial ever found and diamond particles are often used in the field of composite plating to produce the hard and wear resistant coating [47]. In our previous experiment [810], we have demonstrated that the anti-wear performance of a Ni/diamond composite coating depends on the volume percentage of the diamond particles in the nickel matrix and the coating with high volume percent of the reinforcement exhibited the better anti-wear performance. In order to reinforce high amount of particles into a nickel matrix, we have been using redox-active surfactants (AZTAB, Fig. 1) containing an azobenzene moiety [810]. Using a cationic type of azobenzene surfactant, we were able to co-deposit up to about 45 vol.% of diamond particles ( = 1.5 m) with nickel.

N.K. Shrestha et al. / Diamond & Related Materials 15 (2006) 15701575

1571

H 5C2

C2H4 N

(CH3)3, Br

Fig. 1. Molecular structure of AZTAB (4-Ethylazobenzene 4-(Oxyethyl) trimethylammonium Bromide).

The volume percent of the co-deposited particles depends on the interactions among AZTAB, particles and the electrode. The degree of this interaction can be varied if particle size is altered and thereby, the rate of particle co-deposition may be altered. On the other hand, by varying the size of the reinforced particles and their volume fraction, the mechanical properties like hardness and wear of the coating may be varied substantially through changes in the microstructure, the morphology, and the nature of the interface between matrix and reinforcement. Therefore, the present investigation was carried out to study the effect of particle size of the diamond on the co-deposition of the particles with nickel under the influence of a cationic type azobenzene surfactant and to investigate how the degree of hardness and wear resistance of a Ni/diamond coating varies with the particle size. 2. Experimental All chemicals were used without further purification. Altogether five different sizes of diamond particles were used in the present investigation. The size of these diamond particles was labeled by the manufacturer (TOMEI Diamond Co. Ltd., Japan) as 00.25, 0.51.5, 12, 23 and 26 m. The mean sizes of these particles confirmed by a particle size analyzer were 0.21, 1.12, 1.46, 1.91, and 2.91 m, respectively. A copper plate with a dimension of 10 10 0.3 mm was used as the substrate. This substrate was polished and cleaned as described previously [8] and was placed in a vertical position parallel to a nickel anode at a distance of 2 cm in Watt's bath containing 300 g dm 3 NiSO46H2O, 60 g dm 3 NiCl26H2O, and 40 g dm 3 H3BO4. The pH of this bath was adjusted to 1 with a concentrated solution of HCl (35%). A known amount of AZTAB was dissolved in the above bath and then particles were added to the bath. This bath was ultrasonically agitated for 40 min using an ultrasonic disruptor (UR-200P, Tomy Seiko Co., Ltd., Japan) and the mixture was stirred for an hour using a magnetic stirrer. This bath was then subjected to the composite plating at the temperature of 50 C and the particles during plating were kept in suspension using a magnetic stirrer. In all cases, electrodeposition of the composite coating was carried out for 20 min. After plating for 20 min, the deposits were rinsed thoroughly with running water and wiped simultaneously with a wetted tissue paper. Then the deposits were cleaned ultrasonically in water for 20 min. The surface and the crosssectional morphologies of the coatings were studied using a scanning electron microscope (SEM). The thickness of the coatings was measured from the cross-sectional SEM micrograph of the coatings. The amount of diamond in the coating was determined gravimetrically by dissolving the deposits in nitric acid and separating the particles using a membrane filter paper [8]. The amount of nickel was deduced from the differences between the total weight of the composite and the

diamond therein. From these two data, percent of diamond content in the composite was calculated. Vickers's micro-hardness of the coatings was measured using a micro-hardness tester (MHTZ, Matsuzawa Co. Ltd., Japan) under the indentation load of 1.96 N at five different locations of a specimen and the average value of these five measurements was quoted here as the hardness of that coating. Anti-wear performance of the coatings was investigated using an abrasion tester (NUS-ISO3, Sugawara Test Instrument Co. Ltd., Japan) fitted with a SiC abrader. The coating specimen was oscillated to-and-fro against this abrader under the load of 24.5 N for 300 cycles at dry sliding condition of room temperature in air. Before and after the wear test, specimens were ultrasonically cleaned in acetone for 10 min and the average weight loss of three replicas of the dry specimen was quoted here as the wear amount of that material. The degree of wear resistance of the various coatings was considered as the reciprocal of their wear amount. 3. Results and discussion Previously [810], we demonstrated that AZTAB is electrochemically reduced at the cathode surface during the electrolysis of a nickel bath containing AZTAB. This reduction leads to desorb the AZTAB from particle-surface in the vicinity of the cathode, which then leads to deposit the particles on the cathode. The amount of particles incorporated into the nickel deposit depends on the amount of particles deposited to the cathode by the release of AZTAB and the interaction force between these deposited particles and the cathode regarding how strong or how long these particles can remain attached to the cathode surface in order to be embedded into the growing nickel layer. Altering the size of the particles can alter these two factors and thereby altars the amount of particle co-deposition with nickel. To investigate this effect, 5 different sizes of the particles were used in the present study. Based on our previous experiments [810], 10 g dm 3 of these particles was used in the plating bath and the temperature of the bath was kept 50 C. Keeping these two parameters constant in all the experiments of
50
0.21 m 1.12 m 1.46 m 1.91 m 2.91 m

Diamond content/vol.%

40 30 20 10 0 0 5 10

15

Current density/A dm-2


Fig. 2. Effect of current density on the deposition of various particle sizes of diamond with nickel under the influence of AZTAB. 10 g dm 3 diamond particles, 10 g dm 3 AZTAB for 0.21 m particle and 6 g dm 3 for bigger particles.

1572

N.K. Shrestha et al. / Diamond & Related Materials 15 (2006) 15701575

the present study, the effect of current density and the concentration of AZTAB for incorporating the various sizes of diamond particles into the nickel deposit was investigated. First of all, the optimum current density for the maximum particle co-deposition with nickel was investigated. In this experiment, 10 g dm 3 of AZTAB was used in the case of 0.21 m diamond particles, whereas 6 g dm 3 was used for bigger particles. As can be seen in Fig. 2, the optimum current density corresponding to the maximum particle co-deposition was different for different particle sizes. The reasons for this variation will be discussed later. Applying the above optimum current densities for various sizes of diamond particles, next the influence of AZTAB for the maximum particle deposition was investigated. Fig. 3 shows the different optimum AZTAB concentration corresponding to the maximum particle codeposition with nickel for various particle sizes. For the smaller particles, this optimum AZTAB concentration was higher and this concentration decreased with respect to the increment of the particle size. In the present investigation regardless of the particle size, the amount of particles in the plating bath was 10 g dm 3. However, altering the size of the diamond, the total number of the particles and their total surface area for a given mass of the diamond will change, which increases with decreasing the particle size. This leads to the adsorption of higher amount of AZTAB on the smaller particles, whereas the amount of AZTAB adsorbed on bigger particles is comparatively less. As a result, the optimum concentration of AZTAB for maximum particle co-deposition decreased with increasing the particle size. Fig. 3 shows that for the same particle size, the particle co-deposition increases up to the maximum, and then it decreases beyond the optimum AZTAB concentration. The reasons for this increasing and decreasing trend of particle codeposition with respect to the AZTAB concentration can be described on the basis of particle dispersion and adsorption equilibrium of AZTAB. Increasing the AZTAB concentration increases the amount of AZTAB adsorption on the particles. This leads to increase the positive surface charge of the particles as well as the particle dispersion. As a result, number of particles arriving at the cathode will increase which will
0.21 m 1.12 m 1.46 m 1.91 m 2.91 m

60

45

30

15

0 0 4 8 12 16

AZTAB/g dm-3
Fig. 3. Effect of AZTAB on the deposition of various particle sizes of diamond with nickel. 10 g dm 3 diamond particles, peak current densities shown in Fig. 2 were applied.

increase the particle co-deposition. The amount of the particles deposited at the cathode surface depends on the desorption of AZTAB from particle surface. As described earlier [813], free AZTAB is first reduced electrochemically at the cathode surface, which decreases its concentration in the vicinity of the cathode. As a result, adsorption equilibrium of AZTAB near the cathode is broken. In order to satisfy this adsorption equilibrium, desorption of AZTAB from particle surface in that region of the interface takes place. After the reduction, AZTAB looses its dispersion ability due to the loss in hydrophobicity. Therefore, the above particles after desorption of AZTAB cannot be re-dispersed, which leads to deposit on the cathode. However, at high concentration, free AZTAB reduced at the cathode surface is soon substituted from the bulk, which makes difficult to break the adsorption equilibrium of AZTAB. As a result, the desorption of AZTAB from particle surface becomes difficult which makes difficult to deposit the particles on the cathode. Probably, this is the reason for decreasing trend of particle co-deposition at higher amount of AZTAB beyond the optimum concentration. Fig. 3 also reveals that decreasing the particle size could increase the amount of particle co-deposition with nickel. Compared to the larger particle, the interaction between the cathode and the smaller particles might be stronger. On the other hand, owing to the large surface area, the larger particles deposited at the cathode surface might face the greater counter force exerted by the hydrodynamic motion of the electrolyte. Moreover, the larger particles need longer time to be embedded by the growing nickel crystal. All these factors can dislodge the larger particles from the cathode surface before being entrapped into the growing nickel layers whereas the smaller particles being strongly adhered to the cathode cannot be removed as easily as the larger ones. Probably, these are the reasons for incorporation of higher amount of smaller particles into the nickel deposit. The maximum of 57.6 vol.% diamond particles (Fig. 3) was able to co-deposit with nickel under the assistance of AZTAB in the present investigation. Being a cationic surfactant, adsorption of AZTAB develops a net positive charge to the particles' surface, which assists the movement of the particles towards the cathode and favors the particle co-deposition. However, it does not mean that all those particles that arrived at the cathode surface undergo codeposition with the nickel. For the particles to co-deposit with nickel in a bath, these particles must cling to the cathode surface for a long time enough to be entrapped into the growing nickel crystal. In this respect, surfactants, which are used as a dispersing agent in a plating bath, can act an active role. In our previous experiment [8], we have shown that a redox-inactive surfactant adsorbed on the cathode surface acts as a physical barrier for nickel deposition and it also hinders the approaching of the particles due to the repulsive interaction between the surfactant layers on the cathode and the approaching particles. Owing to this phenomenon across the cathodeelectrolyte interface, the amount of particles incorporated into the nickel deposit was very less. On the other hand, we have demonstrated that the azobenzene surfactant (AZTAB) is reduced at the cathode surface during

Diamond content/vol.%

N.K. Shrestha et al. / Diamond & Related Materials 15 (2006) 15701575

1573

the deposition of nickel in composite plating and hence, it does not block the cathode surface as a physical barrier as in the case of a redox-inactive surfactant. Moreover, the reduction of AZTAB leads to desorb the AZTAB from particles near the cathode and this phenomenon leads to deposit these particles at the cathode surface, which can cling

to the electrode surface with sufficient tenacity so that these particles can be embedded into the growing nickel layer. Unlike the AZTAB, the redox-inactive surfactant cannot be desorbed from the particles and hence, the deposition of these particles from their dispersion state is difficult. Even in the case of deposition, the particles loosely adsorb to the cathode

Fig. 4. SEM images of (A) surface of Ni/diamond0.21 m57.6 vol.%, (B) surface of Ni/diamond1.12 m48.7 vol.%, (C) surface of Ni/diamond1.46 m 45.5 vol.%, (D) surface of Ni/diamond1.91 m37.2 vol.%, (E) surface of Ni/diamond2.91 m33.8 vol.%, (F) cross section of c, and (G) EDS mapping of Ni in the coating a at the magnification of 100 times.

1574

N.K. Shrestha et al. / Diamond & Related Materials 15 (2006) 15701575

surface as these particles deposit with the surfactant on their surface. As a result, these particles can be easily re-dispersed into the electrolyte before being embedded into the growing nickel layer. Due to these reasons, the amount of particles incorporated into the nickel deposit from a bath containing AZTAB was considerably higher than that from a bath containing a redox-inactive surfactant [8]. Owing to the larger surface area, the amount of AZTAB adsorbed on 10 g dm 3 of the smaller particles must be higher than that on 10 g dm 3 of the bigger particles. As explained earlier, this adsorbed AZTAB should desorb in order to deposit the particle at the cathode surface and it was believed here that the desorption of AZTAB from particle near the cathode takes place by the reduction of an equivalent amount of free AZTAB. Hence, higher current density is needed to reduce the higher amount of AZTAB in the case of smaller particles. As a result, a decreasing trend of the optimum current density with respect to the increment of the particle size was observed in the present investigation (Fig. 2). All the coatings prepared at the optimum conditions were smooth and uniform. This type of surface morphology was also confirmed by the SEM examination of the surface of the coatings. The SEM images of the coatings with the maximum particle content of the various sizes of the particles are shown in Fig. 4. These SEM images reveal the uniform distribution of the particles in the coating. On the other hand, the crosssectional SEM image (Fig. 4F) also reveals the distribution of the particles throughout the coating and it shows that the coating has a uniform thickness of about 1617 m. The uniformity of the coatings was also confirmed by the uniform distribution of the nickel exhibited by the EDS mapping of the elemental Ni in the composite coating (Fig. 4G). However, some micro-cracks were observed in the case of the coatings with the particle sizes bigger than 1.12 m. The presence of particles causes the distortion of the nickel crystal, which develops the internal strain in the nickel matrix. In the case of a big particle, this distortion is large which leads to develop the large internal strain in the matrix. Therefore, the micro-cracks in presence of bigger particle in

0.21 1000

Micro-hardness/Hv

1.12 1.46 1.91 750 2.91


500

511 426 405 509 379.7

1007.1 934 854 567.9 441

25 vol.% Max. vol.% as shown in Fig. 3

250

0 0 1 2 3

Particle size/m
Fig. 6. Effect of particle size and particle content on the micro-hardness of the Ni/diamond composite coatings.

Micro-hardness/Hv

200 0 0 10

228.2548 362 405 650 854

0 20 30 40 50

Diamond content/vol.%
Fig. 5. Effect of particle content on the micro-hardness and wear of the Ni/ diamond (1.46 m) composite coatings.

Wear amount/mg ...

10000 12.6 17.1 800 26.3 45 600 0 18 400 25 36

8.3 6.4 5.6 1.2 0.7

hardness wear

the coating are probably due to the result of this strain. On the other hand, a large number of micro-cracks were also observed especially when a large amount of AZTAB was used in the bath or when a large current density was applied during the electrolysis of the bath. These micro-cracks might be due to the strain caused by the incorporation of some AZTAB or by the incorporation of hydrogen gas into the deposits. The cross-section morphology of a composite coating (Fig. 4F) shows that the diamond particles were distributed not only on the surface, but they are well distributed throughout the coatings. Such uniformity of the distributed phase is essential for applying the composite coating on tooling materials because for every diamond particle plug out from the matrix, a new particle will appear on the surface of the coating. In this way, the performances of the tools virtually remain unchanged even for prolong use of the tools. The hardness and wear amount of the Ni/diamond coatings reinforced with 1.46 m diamond is shown in Fig. 5. These figures reveal that the hardness as well as the wear resistance of the coatings increases with the increment of the particle content in the coating. Diamond is a hard and brittle phase whereas; the nickel is a ductile phase. Therefore, the hardness of the nickel coating without any reinforcement was very low and its wear amount was very high (Fig. 5). As the incorporation of diamond particle in the coating was increased, the hardness of the coating also increased while the wear amount of the coating decreased showing the better wear resistance of the coating with the increment of the reinforcement. However, there was no linear relationship between the hardness and the wear resistance of the coatings. For example, the micro-hardness in the present investigations gradually increased up to about 25 vol.% of the reinforcement beyond which, it increased sharply, while the amount of wear sharply decreased up to about 25 vol.% of the reinforcement, after which, it decreased more gradually as shown in Fig. 5. Apart from the hardness, the wear resistance of a particle reinforced composite coating also depends on the morphology of the coating especially about how the dispersed particles are anchored in the metallic matrix and how strong is the interface

N.K. Shrestha et al. / Diamond & Related Materials 15 (2006) 15701575

1575

between these particles and the matrix. As revealed by the SEM images (Fig. 4) of the composite coatings, the distribution of incorporated diamond particles was uniform. This type of morphology favors the wear resistance of the coating compared to the presence of agglomerated particles in the deposits. Another factor, which influences the wear of a particle reinforced metal matrix composite coating, is the interfacial strength between the particles and the metallic matrix and this strength probably depends on the particle size. In the present investigation, this interfacial strength could not be measured directly. However, the influence of the particle size on the micro-hardness and the wear were studied. The micro-hardness of the Ni/diamond composite coatings reinforced with the same amount of diamond particles (about 25 vol.%) hardly varied with the particle size (Fig. 6). Like the hardness, the amounts of wear of these coatings were also approximately the same and this amount was almost equivalent to that shown in Fig. 5. In contrast to this result, the micro-hardness of the corresponding coatings with different content of the reinforcement was found to be increased with the increment of the volume percent of reinforcement and the maximum micro-harness of about 1000 Hv was observed in the case of a composite coating containing 57.6 vol % of the diamond particle (Fig. 6). These results suggest that the micro-hardness is controlled by the volume content of the diamond, which is much harder than the nickel matrix. Compared to the nickel coating without any reinforcement, the above coating with 57.6 vol.% of the reinforcement was approximately 4.5 times harder. However, the amounts of wear of the above coatings with different particle sizes and with the corresponding maximum particle content (33.857.6 vol.%, Fig. 3) were not very different. A similar result was described earlier in Fig. 5, in which the wear amount did not vary significantly beyond 25 vol.% of the reinforcement. From these results, it is obvious that the influence of the reinforcement for the wear resistance of a coating is relatively higher in the beginning and as the amount of the reinforcement is increased, the relative contribution of this reinforcement gradually decreases. However, the absolute contribution, though gradually, increases with the increment of the amount of reinforcement. In the present investigation, the wear resistance of the composite coating containing 57.6 vol. % of diamond particles was about 14 times higher than that of the nickel coating without any reinforcement.

4. Conclusion The present investigation revealed that: (a) A composite coating of Ni/diamond with high particle content can be prepared with the assistance of an azobenzene surfactant and the amount of the particle co-deposition can be increased if smaller particles are used in the plating bath. (b) Increasing the volume percent of reinforcement, the micro-hardness and the wear resistance of a particle reinforced Ni/diamond composite coating can be increased. However, no linear relationship between the hardness and the wear existed. (c) The micro-hardness and the wear resistance of a composite coating were hardly influenced by the particle size of the reinforcement. However, these properties were influenced significantly by the volume percent of the reinforcement. The above statements hold true within the particle size of the diamond used in the present investigation. However, a significant influence of the particle size in the co-deposition process and the coating's properties may be observed if very smaller or very bigger particles than that used in the present studies are investigated. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] A. Hovestad, L.J.J. Janssen, J. Appl. Electrochem. 25 (1995) 519. E.P. Rajiv, S.K. Seshadri, Bull. Electrochem. 8 (1992) 376. K. Helle, F. Walsh, Trans. IMF 75 (1997) 53. C.R. Lin, C.T. Kuo, Diamond Relat. Mater. 7 (1998) 903. G. Sheela, M. Pushpavanam, Metal Finish. 1000 (2002) 45. E.C. Lee, J.W. Choi, Surf. Coat. Technol. 148 (2001) 234. V.V.N. Reddy, B. Ramamoorthy, P.K. Nair, Wear 239 (2000) 111. T. Takebe, T. Matsuzaki, N.K. Shrestha, T. Saji, J. Surf. Fin. Soc. Jpn. 54 (2003) 610. N.K. Shrestha, M. Masuko, T. Saji, Wear 154 (2003) 555. N.K. Shrestha, G. Kobayashi, T. Saji, Chem. Lett. 33 (2004) 984. T. Saji, K. Ebata, K. Sugawara, S. Liu, K. Kobayashi, J. Am. Chem. Soc. 116 (1994) 6053. Y. Ito, N. Kashihar, T. Saji, J. Jpn. Colour Mater. 75 (2002) 56. Y. Ito, T. Saji, Langmuir 18 (2002) 6633.

S-ar putea să vă placă și