You are on page 1of 8

Applied Surface Science 258 (2012) 4260–4267

Contents lists available at SciVerse ScienceDirect

Applied Surface Science


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

Effect of particle concentration on the structure and tribological properties of


submicron particle SiC reinforced Ni metal matrix composite (MMC) coatings
produced by electrodeposition
H. Gül b,∗ , F. Kılıç a , M. Uysal a , S. Aslan a , A. Alp a , H. Akbulut a
a
Sakarya University, Engineering Faculty, Department of Metallurgical & Materials Engineering, Esentepe Campus, 54187, Sakarya, Turkey
b
Duzce University, Gumusova Vocational School, Department of Metallurgy, 81850, Duzce, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: In the present work, a nickel sulfate bath containing SiC submicron particles between 100 and 1000 nm
Received 6 June 2011 was used as the plating electrolyte. The aim of this work is to obtain Ni–SiC metal matrix composites
Received in revised form (MMCs) reinforced with submicron particles on steel surfaces with high hardness and wear resistance for
13 December 2011
using in anti-wear applications such as dies, tools and working parts for automobiles and vehicles. The
Accepted 14 December 2011
influence of the SiC content in the electrolyte on particle distribution, microhardness and wear resistance
Available online 8 January 2012
of nano-composite coatings was studied. During the electroplating process, the proper stirring speed was
also determined for sub-micron SiC deposition with Ni matrix. The Ni films were characterized by scan-
Keywords:
Electrocodeposition
ning electron microscopy (SEM) and X-ray diffraction (XRD) analysis. The depositions were controlled to
Nano-SiC obtain a specific thickness (between 50 and 200 ␮m) and volume fraction of the particles in the matrix
Dispersion strengthening (between 0.02 and 0.10). The hardness of the coatings was measured to be 280–571 HV depending on the
Microstructure particle volume in the Ni matrix. The tribological behaviors of the electrodeposited SiC nanocomposite
Friction coatings sliding against an M50 steel ball (Ø 10 mm) were examined on a tribometer. All the friction
Wear resistance and wear tests were performed without lubrication at room temperature and in the ambient air (with
a relative humidity of 55–65%). The results showed that the wear resistance of the nanocomposites was
approximately 2–2.2 times more than those of unreinforced Ni.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction ites. Composite coatings can be prepared by techniques such as


thermal spray, physical vapor deposition, chemical vapor deposi-
Composite electroplating is a method of co-depositing fine tion and electrocodeposition. Because electrocodeposition can be
particles of metallic, non-metallic or polymeric compounds in conducted at normal pressure and ambient temperature with low
the plated layer to improve material properties such as wear cost, high deposition rates and homogenous particle distribution,
resistance, lubrication, or corrosion resistance [1–5]. During the it has been considered one of the most important techniques for
electroplating process, insoluble materials are suspended in a con- producing composites [3,9].
ventional plating electrolyte and captured in the growing metal Ni matrix reinforced with silicon carbide (SiC) have been exten-
film. The second phase material can be hard oxides (Al2 O3 , TiO2 , sively studied and widely used to protect against friction inside
SiO2 ) [1–3], carbides particles (SiC, WC) [1,2,5], diamond, solid parts of cylinders in the automotive industry [15]. Considerable
lubricants (PTFE, graphite, MoS2 ) [1,5], and even microcapsules research has been focused on the impact of the electrodeposition
containing liquids. In general, having second phase particles in a parameters on the electrolytic co-deposition process of SiC with Ni
co-deposited film results in a variety of improvements in film prop- and the properties of the composite coatings. These parameters
erties, such as increased microhardness [5–15], yield strength [5], include electrolysis conditions [1,5,7,15,16] (composition of the
tensile strength, increased wear [4–15] and corrosion resistance electrolytic bath, presence of additives, pH value), current condi-
[4], self-lubrication and high temperature inertness [4], along with tions [1,4,5,7,15] (such as the imposed current and current density
chemical and biological compatibility [4,6–8]. Many preparation values) and properties of the reinforcing particles [1,4,5,7,15,16]
techniques have been investigated to novel metal-based compos- (such as particle size, surface properties, concentration and type of
dispersion in the bath). In general, the amount of embedded SiC
particles has been observed to increase with the concentration of
∗ Corresponding author. Tel.: +90 264 295 57 62; fax: +90 264 295 56 01. suspended SiC particles and the presence of additives in the elec-
E-mail address: harungul@duzce.edu.tr (H. Gül). trolyte [6,7,15,17–24]. From previous reports, decreasing the SiC

0169-4332/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.apsusc.2011.12.069
H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267 4261

particle size can also affect the wear and corrosion resistance in a Table 1
Bath compositions and electrodeposition conditions for MMC reinforced with SiC
positive way. These phenomena have mainly been attributed to the
submicron particles.
hardening of the metal matrix by finely dispersed particles, which
weakened the plowing effect, reduced adhesive wear, retarded the Nickel sulfate (Ni2 SO4 ·6H2 O) (g/L) 300
Nickel chloride (NiCl2 ·6H2 O) (g/L) 50
grain growth in the matrix at elevated temperatures, and thereby
Boric Acid (H3 BO3 ) (g/L) 40
sustained adequate mechanical properties [7,13,15,23]. Thus, par- Sodyumdodecyl sulfate (g/L) 0.2
ticle size plays an important role in the electrocodeposition process Cetyltrimethylammonium bromide (CTAB) (mg/L) 300
and can affect the mechanical properties of the composite coatings Silicon carbide (SiC) (g/L) (0.1–1 ␮m) 5, 10, 20, 30
[7,24]. pH 4
Temperature (◦ C) 45
Generally, the mechanisms of strengthening of alloys and
Current density (A/dm2 ) 3
polycrystalline metals can be described as follows: (I) grain refine- Stirring Speed (rpm) 250, 650
ment strengthening from Hall–Petch relationship; (II) dispersion Plating time (h) 2
strengthening due to Orowan mechanism; (III) solid solution
strengthening and (IV) crystal orientation. The wear resistance and
hardness of composite coatings are related to each other as well diffractometer scanning at a speed of 1◦ /min in the 2 range
as to the particle content of the deposit. With increasing the vol- of 10–100◦ . The hardness of the coatings was measured using
ume fraction of particles in the deposited layer, wear resistance a Vicker’s microhardness indenter (Leica VMHT) with a load of
increases and thus, the weight of the removed material due to wear 50 g. The reciprocating tribological behaviors of the coatings sliding
is decreased [25,26]. against an M50 steel ball (Ø 10 mm) were examined on a Tribome-
The problem of agglomeration can be greatly reduced by adding ter (CSM Instruments) designed according to DIN 50 324 and ASTM
the surfactant cetyltrimethylammonium bromide (CTAB) [15]. G 99-95a in a ball-on-disk configuration. The wear tests were per-
It has been confirmed that suitable surfactants could not only formed at a constant applied load of 1.0 N with a sliding speed of
improve the stability of a suspension by increasing the wettabil- 50 mm/s. The SEM and energy dispersive X-ray spectroscopy (EDS)
ity and the surface charge of suspended particles, but they can also was used to study the morphologies of the wear tracks. The parti-
enhance the electrostatic adsorption of suspended particles on a cle volume fractions were calculated directly from the 6060 LV SEM
cathode surface by increasing their positive charge [15,16]. Hou image analysis program, which was based on phase area method.
et al. [15] demonstrated by the dynamic light scattering method
that inert SiC particles formed large agglomerates in the plat- 3. Results and discussions
ing bath. When CTAB was added, the SiC particle agglomerates
decreased, and a high percentage of uniformly distributed, sub- 3.1. Effects of the stirring speed on co-deposited SiC particles
micron SiC particles were deposited. Increasing the amount of
surfactants adsorbed on the surface of the inert particles would In this study, two different stirring speeds initially were applied
enhance the possibility of the weak and strong adsorption effect to observe the effect of stirring speed in the course of electrode-
[15]. position. Hence, it was chosen one low speed (250 rpm) and one
The aim of this work is to deposit Ni-MMCs reinforced with high speed (650 rpm). Fig. 1 shows SEM micrographs of coatings
submicron-sized SiC particles on steel surfaces to improve surface which performed to observe the effect of stirring speed on the
hardness and wear resistance. Introducing high amount of sub- deposited SiC vol.% in the electroplated Ni matrix. The volume frac-
micron ceramic particles is one of the main targets to increase tion of co-deposited SiC was 2.54 vol.% increasing the stirring speed
wear resistance of Ni based coatings and in general, it is difficult to 650 rpm increased the amount of co-deposited SiC to 10.05 vol.%.
to deposit sub-micron particle over 10 vol.% content by electrode- Because stirring at 650 rpm produced layers with a large amount of
position, since poor wettability causes a detrimental effects for SiC particles without segregation, the stirring speed for all coatings
the properties. Therefore, the process parameters of stirring speed was fixed at this level. It is known that in micron-sized parti-
and particle content in the electrolyte were extensively studied cle reinforced composite coatings vigorous stirring is necessary to
to deposit high amount of sub-micron particles by using a con- obtain homogenous distribution [27]. This effect also supported in
stant amount of surfactant. In this work, it was aimed to deposit our present work (Fig. 1). It is also best known decreasing second
as high as 10 vol.% SiC and systematically to investigate the effect phase particle size leads to increase yield strength and hardness
of the particle volume fractions on the wear performances of the of the resultant composite whereas, coarse particles improve wear
electrodeposited Ni coatings. resistance. In this work, it was also aimed to obtain hard and wear
resistant Ni matrix coatings and therefore, high stirring speed was
chosen.
2. Experimental procedure
3.2. Effect of particle concentrations in electrolyte on
The plating electrolyte used in the fabrication process was a deposition/deposits
nickel sulfate bath. The composition and the range of experimen-
tal process parameters are shown in Table 1. The range of SiC The effect of SiC concentration in the electrolyte on the content
particle size used in the experiment was 0.1–1 ␮m. A Ni plate of of co-deposited SiC can be observed in Fig. 2. The volume fraction of
30 mm × 35 mm was used as an anode, and a stationary iron sub- co-deposited SiC appeared to increase with the SiC concentration
strate was used as a cathode. The bath was heated to 45 ◦ C with a in the electrolyte (Fig. 3a). The volume fraction of the co-deposited
fixed pH of 4 and stirred with a magnetic stirrer at a stirring speed SiC can be increased up to approximately 10 vol.% by increasing the
value of 250 and 650 rpm, respectively. The amount of surfactant, concentration of SiC in the electrolyte to 20 g/L. The experimental
CTAB was fixed at 300 mg/L, since agglomeration free sub-micron data shown in Fig. 3a revealed the relationship between the concen-
SiC particle depositing was provided in this concentration. tration of SiC in the electrolyte and dispersed SiC particle content
After co-deposition, a scanning electron microscope (SEM, in the deposited layer. It is clear that the amount of co-deposited
model JEOL – JSM 6060 LV) was used to observe the surface and SiC increased with the SiC content in the electroplating bath up to
the cross-sectional microstructures of the deposits. XRD analy- a certain SiC concentration. The amount of co-deposited SiC par-
sis was carried out with a Rigaku D/MAX/2200/PC model X-ray ticles was reached to a maximum value at 20 g/L particle content
4262 H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267

Fig. 1. Cross-sectional SEM micrographs of the composite coatings, showing the distribution of SiC particles deposited at the stirring speeds of (a) 250 rpm and (b) 650 rpm.

Fig. 2. Cross-sectional SEM micrographs of SiC particle distributions in the composite coatings. The electrolyte used to prepare the coatings has SiC concentrations of (a)
5 g/L, (b) 10 g/L, (c) 20 g/L and (d) 30 g/L.

a 12 b
o Nickel
Volume fraction of SiC (vol. %)

SiC
10
(111)

(200)

(311)
(220)

(222)

o o o
Ni o o
8
Intensity

5 g/l SiC
6
10 g/l SiC
4
20 g/l SiC

2
30 g/l SiC

0
5 10 20 30 20 40 60 80 100

Particle concentration in electrolyte (g/l) 2θ

Fig. 3. (a) The volume fraction of SiC particles deposited at various SiC concentrations in the electrolyte, (b) effect of various SiC concentrations in the electrolyte on XRD
patterns.
H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267 4263

in the electrolyte and then decreased when the concentration was


further increased to 30 g/L. 600 10,05% 9,32%
Lee et al. [24] showed that the maximum amount of co-
deposited SiC corresponded to steady-state equilibrium, where the 500
number of co-deposited SiC particles equals the number of SiC

Microhardness (HV)
particles approaching the cathode surface. When the SiC content 400 4,54%
4,1%
exceeded the maximum amount of co-deposited of SiC at equilib-
rium, suspended SiC particles agglomerated in the plating bath, and
300
the amount of particles incorporated into the Ni matrix decreased.
In addition, the agglomerates themselves can resist being incor-
200
porated into the Ni matrix when the SiC content in the plating
bath is high, thereby shielding the surface of the cathode from
the flux of incoming SiC particles, as reported by Abdel et al. [25]. 100
The decrease in particle content in the electrolyte beyond 20 g/L
was taken place because of two reasons: (i) it is believed that a 0
SiC particle concentration above 20 g/L in the electrolyte can lead Ni 5 10 20 30
to a shielding effect for the particles co-deposited in the Ni layer. Particle concentration in electrolyte (mg/l)
During the co-deposition of the Ni and SiC, the SiC particles were
Fig. 4. Effect of particle concentration in the electrolyte on microhardness of the
suspended in the electrolyte. Not all the SiC submicron particles
coating.
were expected to tend to the cathode and be deposited in the Ni
layer. Because of the elastic collision of the SiC particles inside the
electrolyte, some of the SiC particles were backscattered from the be suggested that introducing different amount of SiC particles into
interface between electrolyte and substrate where the particle den- Ni matrix resulted to change nucleation and growth kinetics of
sity is highest. If a high amount of SiC particles were present in the Ni matrix deposited. The change is expected to cause lattice
the electrolyte, they could backscatter at the cathode–electrolyte distortion and hence, increment in the matrix hardness. Increas-
interface, since the SiC particles were directed to the cathode sub- ing particle content in the electrolyte and thus, in the deposited
strate. (ii) The decrease can take place according to the Guglielmi layer promoted the intensity of the reflections of (2 0 0) and (3 1 1).
two-step adsorption model [28]. Increasing particle concentration Accordingly, the Ni–SiC composite exhibited increased (2 2 2) and
in the electrolyte increases the adsorption, thus resulting in a higher (3 1 1) diffraction planes and attenuated (2 0 0) plane with increas-
volume percent of SiC submicron particles in composite coatings, ing SiC content. As stated in Fig. 2, increasing particle content in the
while the decrease in the volume percent of the SiC submicron par- electrolyte resulted in increasing the co-deposited SiC. However,
ticles at a concentration of SiC particles above 20 g/L is attributed to this effect is not clear in Fig. 3b. The intensity of SiC peak for the
the agglomeration of the SiC submicron particles in the electrolyte composite deposited at 5.0 g/L particle content in the electrolyte
owing to their poor wettability. is 442.0 and gradually increased by increasing particle content in
The amount of co-deposited SiC particles increased significantly the electrolyte and reached to 782.0 for the composite deposited at
up to a SiC concentration of 20 g/L, likely because more particles 30 g/L SiC.
were present in the plating bath. Only particles that clung to the
cathode surface for a sufficient period of time can be successfully
3.3. Effect of particle concentration on microhardness
incorporated into the Ni matrix. Therefore, the entrapment of the
SiC particles into the growing Ni matrix depended on both the rate
The hardness values of the unreinforced Ni coating and the
of SiC particles approaching the cathode surface and the rate of Ni
Ni-MMCs co-deposited with submicron-SiC are shown in Fig. 4.
deposition. Although the number of SiC particles approaching the
The improved hardness due to the increased SiC content in the
cathode surface increased with the SiC content in the plating bath,
deposited layer was highly correlated to the volume fraction of
the capturing capacity of the growing metal remained virtually the
SiC. The microhardness of the Ni–SiC composite coating containing
same [24]. Consequently, only a limited amount of SiC particles was
10.05 vol.% SiC was 571 HV, whereas the microhardness of a pure Ni
co-deposited when the concentration was beyond 20 g/L.
coating was only 280 HV. The enhanced hardness of the composite
Fig. 3b shows the XRD patterns of pure Ni and Ni–SiC nanocom-
can be attributed to the contribution from the soft matrix and to
posites. The (2 0 0) diffraction peak of Ni in the nanocomposite
some extent, from the SiC particulates. The relation between hard-
coating has a lower peak intensity and broader peak width than that
ness and crystallite size is in accordance with the Hall–Petch effect
of the Ni coating (see Fig. 3b) as also evidenced by many researchers
[29].
[28,29]. This was attributed to the decrease in the grain size of the
Ni–SiC nanocomposite coating with the addition of SiC submicron
particles in the plating bath. The submicron particles provided more 3.4. Lattice distortion of nickel matrix
nucleation sites and retarded crystal growth, resulting in a smaller
crystal size for the Ni matrix in the composite coating [17]. Lattice distortion of the Ni matrix for unreinforced alloy and
Pure Ni exhibited two intense reflection peaks of (2 0 0) and Ni–SiC composite matrix was determined by calculating their lat-
(1 1 1), introducing SiC particles led to change the predominant tice constants using basic reflections of the crystal planes. For
reflection peak to (1 1 1). This means mixed (2 0 0) and (1 1 1) planes comparison, the lattice constant for unreinforced Ni coatings was
changed to predominant diffraction in (1 1 1) plane. Srivastava et al. also calculated and compared with co-deposited Ni matrix reflec-
[29] reported same results from sulfamate bath. The reason was tions. The deviation in the lattice constants of the unreinforced Ni
attributed to change surface morphology of the Ni grains. They coatings and composite matrices was assessed as the lattice dis-
reported that the incorporation of nano-SiC has altered the Ni–Co tortion. For a fcc system, the lattice constant can be calculated as
matrix from a dense spherical smooth surface to rough nodular follows [30]:
morphology. As stated by Gyftou et al. [7], the preferred orienta-  
2 2
tion and texture of the Ni growth were also influenced by changing a = (h2 + k2 + l2 ) (1)
duty cycle percent at constant applied current density. Here, it can 4 sin2 
4264 H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267

0,5 also showed a large amount of plastic deformations and seizure of


Ni on the sliding surfaces. Depending on the particle content in the
0,4
deposited layer the amount of wear decreased by 170% (Fig. 7a).
The wear resistances of MMCs are strongly related to the
particle content in the metallic matrix. Fig. 7a shows the relation-
0,3 ship between the particle concentration in the electrolyte, which
depended on the volume fraction of SiC and wear rate. The results
% Distortion

0,2 indicate that the wear resistance increased with the volume frac-
tion of SiC in the deposited layer. The wear rates of composite layers
depended mainly on the particle concentration in the electrolyte.
0,1
The composite that exhibited the highest wear resistance has the
maximum volume fraction of SiC (10.05 vol.%). Several authors
0,0 (111) [13,15,28,31,32] described that the addition of hard microceramic
(200) particles into the metal matrix can improve its hardness and wear
(311)
resistance. These phenomena are mainly attributed to the harden-
0 5 10 15 20 25 30 35 ing of the metal matrix by finely dispersed ceramic particles. These
SiC in electrolyte (g/l) dispersed ceramic particles weaken the plowing effect and adhe-
sive wear, and they retard the grain growth in the matrix at elevated
Fig. 5. Effect of particle concentration in the electrolyte on the lattice distortion of temperatures to sustain good mechanical properties. However, as
the Ni matrix. reported by several authors the particles need to be smaller than
1 ␮m in diameter and uniformly dispersed to exhibit the dispersion
The deviation in the lattice constant depending on the particle hardening effect [25,26,33].
content and deposition parameters was calculated with the follow- The amount and size of the particles define two kinds of
ing equation: reinforcing mechanisms in MMC materials, namely dispersion
 (a − a )  strengthening and particle strengthening. When the dispersed par-
0 1
distortion% = 100 (2) ticle size is higher than 1 ␮m the strength of the metallic matrix
a0
is increased predominantly due to load carrying capacity of the
where a0 is the lattice constant of the unreinforced Ni coating, and reinforcement phase. A dispersion strengthened composite is char-
a1 is the lattice constant of the Ni matrix for composites produced acterized by a dispersion of fine particles with a particle diameter
at different conditions. ranging from 0.01 to 1 ␮m through a dislocation–particle inter-
Fig. 5 shows the relationship between lattice distortion and action or an Orowan hardening mechanism [32]. In this case,
particle content in the electrolyte. The initial introduction of SiC the matrix carries the load, and the fine particles impede the
particles at a concentration of 5 g/L into the electrolyte led to a motion of dislocations [34]. Thus, for the Ni–SiC nanocomposite
negative lattice distortion at all crystal planes. However, a parti- in this experiment, the enhancement was related to the disper-
cle concentration of higher than 5 g/L in the electrolyte gradually sion strengthening effect caused by SiC particles that impeded
increased lattice distortion. The reason for the initial negative and the motion of dislocations in the metallic matrix of the compos-
subsequent positive lattice distortion with increasing particle con- ite coatings [15,35]. During the friction process, the co-deposited
tent in the electrolyte is not yet clear. It is thought that up to SiC submicron particles that carried the loads transferred from
10 g/L SiC addition into the electrolyte is not effective to change the matrix gradually protruded out of the matrix. As a result, the
the growth structure of the Ni matrix. As stated in XRD results amount of thermal plastic deformation and scuffing wear in the
increasing particle content result in changing predominant growth sliding was reduced. The protrusion of the SiC particles was also
planes. However, there should be some other reasons for this pos- the reason why the friction coefficient of the Ni–SiC nanocomposite
itive lattice distortion related with the surfactant additions. As was more than the as-deposited Ni film.
known surfactant CTAB is a hydrocarbon type chemical compound. In the alloys that reinforced with submicron particles (smaller
Another possible reason is thought that during electrodeposition, than 1 ␮m) another possible effect are the differences thermal mis-
the surfactant was ionized for constituting a negative charge at the match between constituents, and matrix grains size refinement. As
particle surfaces. C–Ni equilibrium diagram showed a 1.6 wt.% sol- increasing in the second phase content the grain size was reduced
ubility of C at 1025 ◦ C but the diagram does not give information if of the Ni matrix was reported [36]. Therefore, the increase in the
Ni solves C at 45 ◦ C at which the coating processes were carried out. microhardness values of the Ni–SiC nanocomposites may be due to
It is supposed that the ionized surfactant leads the C ions diffuse smaller grain size of the Ni matrix favored by the submicron-sized
into Ni matrix and behave as an interstitial element and so causes SiC particles.
a positive distortion in the Ni lattice [1]. In the deposited layer with the 5 g/L particle concentration,
the wear rate was approximately 12 × 10−4 mm3 /Nm. However,
3.5. Wear and friction the wear rate decreased to 5.5 × 10−4 mm3 /Nm when the parti-
cle concentration in the electrolyte increased to 20 g/L, indicating
3.5.1. Effect of particle concentration on the wear and friction approximately a 2.2-fold increase in the wear resistance. This is
properties because of the increased amount of SiC particles resulted in increas-
Fig. 6 shows the SEM images of the worn surfaces at a low ing matrix strength and also load carrying capacity of SiC particles,
magnification that indicates the effect of particle concentration especially the particles that their sizes are close to 1 ␮m.
in the electrolyte. As evidenced in the figure, the amount of wear The friction coefficient of pure Ni is 0.2. Introducing 5 g/L and
decreased in the deposited layer when the SiC particle content in 10 g/L SiCp into the electrolyte resulted in an effective increase
the electrolyte increased up to 20 g/L because the width of the wear in the friction coefficient of the Ni matrix. Introducing SiCp into
tracks decreased. From the low magnification micrographs, a high Ni resulted in increasing friction coefficient. Introducing 5 g/L SiCp
level of plastic deformation was seen on the surface of the com- into the electrolyte increased friction coefficient to 1.22. As seen
posites produced with a 5 g/L and 10 g/L particle content in the in Fig. 7b, the friction coefficient increased to 1.38 for the com-
electrolyte. The composite deposited at a particle content of 30 g/L posite produced with a particle concentration of 10 g/L, which
H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267 4265

Fig. 6. SEM morphologies of the wear tracks of composite coatings prepared with different particle concentrations in the electrolyte, shown at low magnifications: (a) 5 g/L
SiC (4.1 vol.% SiC), (b) 10 g/L SiC (4.54 vol.% SiC), (c) 20 g/L SiC (10.05 vol.% SiC) and (d) 30 g/L SiC (9.32 vol.% SiC).

subsequently decreased as the particle concentration exceeded coating surfaces. At the particle concentration of 20 g/L, the worn
10 g/L in the electrolyte. The reason of this increment in the fric- surfaces were fairly smooth. However, the worn surfaces demon-
tion coefficient with increasing particle content in the electrolyte strated a predominantly abrasive mechanism when the particle
from 5 g/L to 10 g/L is not so clear. The reason of this increment concentration was 30 g/L. Increasing particle content in the elec-
is believed to be caused from insufficient load carrying effect of trolyte and thus, increasing particle content in the electrodeposited
SiC particles. High interparticle spacing between co-deposited SiC layer resulted in decreasing amount of the surface cracks which
particles is thought was not enough to create effective dispersion shows the wear takes place also by abrasive type of mechanisms.
hardening. During the sliding process, the stress is concentrated Formations of the surface cracks are main evidence of the wear
between the Ni and SiC particles and resulted in particle pull out, product smearing on the worn surface and then subsequent delam-
which deteriorated friction coefficient due to abrasive effect. ination type of the wear.
Fig. 8 shows the SEM morphologies of worn surfaces at high EDS analysis of the coatings showed formation of an inhomoge-
magnifications, demonstrating different wear mechanisms on the neous transfer layer as the SiC particle content increased; however,
composites. At low particle concentrations, the worn surfaces the formation of this layer was dependent on the CTAB and SiC
showed seizure of the wear debris on the sliding surface and particle content in the electrolyte. From the EDS spectra, it was
subsequent plastic deformation hardening evidenced by microc- concluded that increasing the particle content led to the forma-
racks on the surface. Increasing the charge concentration resulted tion of a transfer layer that consisted of a mixture of iron debris,
in fewer deformation hardenings. The wear mechanism transited which comes from counterpart steel ball, Ni matrix and fragments
from adhesive to abrasive, as evidenced by small grooves on the of second phases on the sliding surfaces.

a 20 b 1,6
19
Wear Rate ( x10-4 mm3/Nm)

18
17 1,4
Coefficient of Friction (μ)

16
15 1,2
14 Ni 4,1%
13
12 4,54% 1,0
11
10
9 0,8
8 9,32%
7 0,6
6
5 10,05%
4 0,4
3
2 Ni
1 0,2
0
-5 0 5 10 15 20 25 30 35 -5 0 5 10 15 20 25 30 35

SiC concentration in electrolyte (g/l) SiC concentration in electrolyte (g/l)

Fig. 7. Effect of the particle concentration in the electrolyte on: (a) wear rate and (b) the friction coefficient of the Ni–SiC composite coatings.
4266 H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267

Fig. 8. SEM morphologies of the wear tracks of composite coatings prepared with different particle concentrations in the electrolyte, shown at high magnifications: (a) 5 g/L
SiC (4.1 vol.% SiC), (b) 10 g/L SiC (4.54 vol.% SiC), (c) 20 g/L SiC (10.05 vol.% SiC) and (d) 30 g/L SiC (9.32 vol.% SiC).

As seen from the above results, introducing SiC submicron parti- Acknowledgments
cles (between the sizes of 100–1000 nm) in the matrix can improve
the hardness and wear performance of Ni–SiC nanocomposite This work was supported by the Scientific and Technical
coatings. Increasing surface hardness and therefore, poor surface Research Council of Turkey (TUBITAK) under the contract number
deformation ability is arising from the hardness of the resultant 106M253. The authors thank the TUBITAK MAG workers for their
composite. financial support.

4. Conclusions References

1. Ni MMCs reinforced by submicron-sized (100–1000 nm) SiC par- [1] H. Gül, F. Kılıc, S. Aslan, A. Alp, H. Akbulut, Characteristics of electro-co-
ticle were successfully co-deposited by D.C. electroplating with deposited Ni–Al2 O3 nano-particle reinforced metal matrix composite (MMC)
coatings, Wear 267 (2009) 976–990.
up to 10.05 vol.% particle concentration. [2] I. Haq, T.I. Khan, Tribological behavior of electrodeposited Ni–SnO2 nanocom-
2. The electrodeposited coatings reinforced with submicron-SiC posite coatings on steel, Surf. Coat. Technol. 205 (2011) 2871–2875.
particles yielded hardness values as high as 571 HV because of [3] P. Baghery, M. Farzam, A.B. Mousavi, M. Hosseini, Ni–TiO2 nanocomposite coat-
ing with high resistance to corrosion and wear, Surf. Coat. Technol. 204 (2010)
unique dispersion effect, whereas the hardness value of the unre- 3804–3810.
inforced Ni matrix hardness was 280 HV. [4] Y.J. Xue, H.B. Liu, M.M. Lan, J.S. Li, H. Li, Effect of different electrodeposition
3. Increasing the particle content in the electrolyte resulted in an methods on oxidation resistance of Ni–CeO2 nanocomposite coating, Surf. Coat.
Technol. 204 (2010) 3539–3545.
increase of particle concentration of the co-deposited coatings,
[5] E. Rudnik, L. Burzynska, Ł. Dolasinski, M. Misiak, Electrodeposition of nickel/SiC
up to a maximum of 20 g/L. Further increase in the particle con- composites in the presence of cetyltrimethylammonium bromide, Appl. Surf.
tent in the electrolyte decreased the particle concentration in Sci. 256 (2010) 7414–7420.
[6] M.D. Ger, Electrochemical deposition of nickel/SiC composites in the presence
the coating.
of surfactants, Mater. Chem. Phys. 87 (2004) 67–74.
4. Increasing the particle content in the electrolyte resulted in the [7] P. Gyftou, E.A. Pavlatou, N. Spyrellis, Effect of pulse electrodeposition param-
lattice distortion of the Ni matrix. eters on the properties of Ni/nano-SiC composites, Appl. Surf. Sci. 254 (2008)
5. Wear resistance of the coatings increased with SiC particle 5910–5916.
[8] M. Surender, R. Balasubramaniam, B. Basu, Electrochemical behavior of elec-
content up to 20 g/L in the electrolyte. However, the wear resis- trodeposited Ni–WC composite coatings, Surf. Coat. Technol. 187 (2004) 93–97.
tance began to decrease beyond this concentration. Pure Ni [9] L. Shi, C. Sun, W. Liu, Electrodeposited nickel–cobalt composite coating con-
showed adhesive type of the wear mechanisms whereas Ni–SiCp taining MoS2 , Appl. Surf. Sci. 254 (21) (2008) 6880–6885.
[10] L. Shi, C.F. Sun, F. Zhou, W.M. Liu, Electrodeposited nickel–cobalt composite
composites exhibited an abrasive type of wear which its predom- coating containing nano-sized Si3 N4 , Mater. Sci. Eng. A 397 (2005) 190–194.
inance increases with increasing SiCp. [11] L. Burzyńska, E. Rudnik, J. Koza, L. Błaż, W. Szymański, Electrodeposition and
6. The co-deposited Ni–SiC nanocomposite coatings showed higher heat treatment of nickel/silicon carbide composites, Surf. Coat. Technol. 202
(2008) 2545–2556.
friction coefficients and better wear resistance compared to [12] Y. Yao, S. Yao, L. Zhang, H. Wang, Electrodeposition and mechanical and cor-
the as-deposited Ni film. This was attributed to the incorpora- rosion resistance properties of Ni–W/SiC nanocomposite coatings, Mater. Lett.
tion of submicron-sized SiC particles in the deposit that greatly 61 (2007) 67–70.
[13] I. Garcia, J. Fransaer, J.P. Celis, Electrodeposition and sliding wear resistance of
increased the hardness of the composite coating through grain
nickel composite coatings containing micron and submicron SiC particles, Surf.
refinement strengthening and dispersion strengthening. Coat. Technol. 148 (2001) 171–178.
H. Gül et al. / Applied Surface Science 258 (2012) 4260–4267 4267

[14] E. Pompei, L. Magagnin, N. Lecis, P.L. Cavallotti, Electrodeposition of nickel–BN [26] S. Mohajeri, A. Dolati, S. Rezagholibeiki, Electrodeposition of Ni/WC
composite coatings, Electrochim. Acta 54 (2009) 2571–2574. nano composite in sulfate solution, Mater. Chem. Phys. 129 (2011)
[15] K.H. Hou, M.D. Ger, L.M. Wang, S.T. Ke, The wear behaviour of electro- 746–750.
codeposited Ni–SiC composites, Wear 253 (2002) 994–1003. [27] G. Maurin, A. Lavanant, Electrodeposition of nickel/silicon carbide compos-
[16] A. Hovestad, L.J.J. Janssen, Electrochemical codeposition of inert particles in a ite coatings on a rotating disc electrode, J. Appl. Electrochem. 25 (1995)
metallic matrix, J. Appl. Electrochem. 25 (1995) 519–527. 1113–1121.
[17] Q. Feng, T. Li, H. Yue, K. Qi, F. Bai, J. Jin, Preparation and characterization of nickel [28] M.R. Vaezi, S.K. Sadrnezhaad, L. Nikzad, Electrodeposition of Ni–SiC nano-
nano-Al2 O3 composite coatings by sediment co-deposition, Appl. Surf. Sci. 254 composite coatings and evaluation of wear and corrosion resistance and
(2008) 2262–2268. electroplating characteristics, Colloids Surf. A: Physicochem. Eng. Aspects 315
[18] S.L. Kuo, Y.C. Chenb, M.D. Ger, W.H. Hwu, Nano-particles dispersion effect on (2008) 176–182.
Ni/Al2 O3 composite coatings, Mater. Chem. Phys. 86 (2004) 5–10. [29] M. Srivastava, V.K. William Grips, K.S. Rajam, Electrochemical deposition and
[19] S.C. Wang, W.C.J. Wei, Kinetics of electroplating process of nano-sized ceramic tribological behaviour of Ni and Ni–Co metal matrix composites with SiC nano-
particle/Ni composite, Mater. Chem. Phys. 78 (2003) 574–580. particles, Appl. Surf. Sci. 253 (2007) 3814–3824.
[20] C.T.J. Low, R.G.A. Wills, F.C. Walsh, Electrodeposition of composite coatings con- [30] I. Misbah-Ul, K.A. Hashmi, M.U. Rana, T. Abbas, Measurement of exchange inter-
taining submicron particles in a metal deposit, Surf. Coat. Technol. 201 (2006) action in Ti-substituted Ni-ferrites, Solid State Commun. 121 (2002) 51–54.
371–383. [31] M.C. Choua, M.D. Ger, S.T. Ke, Y.R. Huang, S.T. Wu, The Ni–P–SiC composite
[21] L. Orlovskaja, N. Periene, M. Kurtinaitiene, S. Surviliene, Ni–SiC composite produced by electro-codeposition, Mater. Chem. Phys. 92 (2005) 146–151.
plated under a modulated current, Surf. Coat. Technol. 111 (1999) 234–239. [32] Y. Zhou, H. Zhang, B. Qian, Friction and wear properties of the co-deposited
[22] A.F. Zimmerman, G. Palumbo, K.T. Aust, U. Erb, Mechanical properties of nickel Ni–SiC nanocomposite coatings, Appl. Surf. Sci. 253 (2007) 8335–8339.
silicon carbide nanocomposites, Mater. Sci. Eng. A 328 (2002) 137–146. [33] G. Wu, N. Li, D. Zhou, K. Mitsuo, Electrodeposited Co–Ni–Al2 O3 composite coat-
[23] I. Garcia, A. Conde, G. Langelaan, J. Fransaer, J.P. Celis, Improved corrosion ings, Surf. Coat. Technol. 176 (2004) 157–164.
resistance through microstructural modifications induced by codepositing SiC- [34] B. Muller, H. Ferkel, Al2 O3 -nanoparticle distribution in plated nickel composite
particles with electrolytic nickel, Corros. Sci. 45 (2003) 1173–1189. films, Nanostruct. Mater. 10 (1998) 1285.
[24] H.K. Lee, H.Y. Lee, J.M. Jeon, Codeposition of micro- and nano-sized SiC particles [35] Y.-J. Xue, X.-Z. Jia, Y.-W. Zhou, W. Ma, J.-S. Li, Tribological performance of
in the nickel matrix composite coatings obtained by electroplating, Surf. Coat. Ni–CeO2 composite coatings by electrodeposition, Surf. Coat. Technol. 200
Technol. 201 (2007) 4711–4717. (2006) 5677–5681.
[25] A. Abdel Aal, K.M. Ibrahim, Z. Abdel Hamid, Enhancement of wear resistance of [36] N.S. Qu, D. Zhu, K.C. Chan, Fabrication of Ni–CeO2 nanocomposite by electrode-
ductile cast iron by Ni–SiC composite coating, Wear 260 (2006) 1070–1075. position, Scripta Mater. 54 (2006) 1421–1425.

You might also like