effect of CTAB- colloids and surfaces - yarbis

Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
Contents lists available at SciVerse ScienceDirect
Colloids and Surfaces A: Physicochemical and
Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa
Effect of CTAB concentration in the electrolyte on the tribological properties of
nanoparticle SiC reinforced Ni metal matrix composite (MMC) coatings produced
by electrodeposition
F. Kılıc¸ a , H. Gül b,∗ , S. Aslan a , A. Alp a , H. Akbulut a
Sakarya University, Engineering Faculty, Department of Metallurgical & Materials Engineering, Esentepe Campus, 54187, Sakarya, Turkey
Duzce University, Gumusova Vocational School, Department of Metallurgy, 81850, Duzce, Turkey
h i g h l i g h t s
We have aimed to incorporate large
amount of SiC into Ni layer with a
plating process.
A nickel sulphate bath the plating
electrolyte contained SiC nanoparticles was used.
The effect of CTAB content on physical and mechanical properties has
been studied.
We have carried out a systematic
study to increase the amount of the
SiC particles.
Subsequently,
investigating
microstructural and wear properties.
a r t i c l e
i n f o
Article history:
Received 13 July 2012
Received in revised form 1 November 2012
Accepted 24 November 2012
Available online 5 December 2012
Keywords:
Co-electrodeposition
Dispersion strengthening
Friction
Wear resistance
g r a p h i c a l
a b s t r a c t
30
20
10
0
-10
-20
-30
0
100
200
300
400
Volume percentage of SiC (vol. %)
b
Zeta potential (mV)
a
Concentration of surfactant (CTAB) (mg/l)
12
10
8
6
4
2
0
0
100
200
300
400
CTAB concentration (mg/l)
a b s t r a c t
In this study, a nickel sulfate bath containing SiC nanoparticles (between 100 and 1000 nm) was used
to obtain hard and wear-resistant nanoparticle reinforced Ni SiC MMCs on steel surfaces for anti-wear
applications, such as dies, tools and working parts. The influence of stirring speed and surfactant concentration on particle distribution, microhardness and wear resistance of nano-composite coatings has been
studied. The nickel films were characterized by scanning electron microscopy (SEM) and X-ray diffraction (XRD). The depositions were controlled to obtain a specific thickness (between 50 and 200 ␮m) and
particle volume fraction in the matrix (between 0.02 and 0.12). The hardness of the resulting coatings
was also measured and found to be 280–571 Hv, depending on the particle volume in the Ni matrix. The
effects of the surfactant on the zeta potential, co-deposition and distribution of SiC particles in the nickel
matrix, as well as the tribological properties of composite coatings, were investigated. The tribological
behaviors of the electrodeposited SiC nano composite coatings sliding against M50 steel ball (Ø 10 mm)
were examined on a CSM Instrument. All friction and wear tests were performed without lubrication at
room temperature and in the ambient air (relative humidity 55–65 %). The results showed that the wear
resistance of the nano composites was approximately 2–2.2 times higher than unreinforced Ni deposited
material.
© 2012 Elsevier B.V. All rights reserved.
1. Introduction
∗ Corresponding author. Tel.: +90 264 295 57 62; fax: +90 264 295 56 01.
E-mail address: [email protected] (H. Gül).
0927-7757/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfa.2012.11.048
Electroplating is a method of co-depositing micron- or nanosized particles of metallic or non-metallic compounds and
polymers with a metal or alloy matrix. Composite deposits are
54
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
used in fields ranging from high-tech industries, such as electronic
components and computers, to more traditional industries such
as the general mechanics and automobile, paper mill, textiles and
food industries. Recently, the main work in this field has focused
almost entirely on the production of wear- and corrosion-resistant
coatings, self-lubricating systems and dispersion-strengthened
coatings [1–3].
Nickel is widely used as a metal matrix. Ni–SiC composites have
been commercialized for the protection of friction parts, combustion engines and casting molds. The co-electrodeposition between
metal Ni and solid particles have been studied extensively. One
interesting phenomenon of this process is that the results are different if the solid particles are different. During co-electrodeposition,
solid insoluble materials are suspended in a conventional plating electrolyte and captured in the growing metal film. For nickel
matrix electrodeposits in particular, a great variety of particles
has been used, such as hard oxides (Al2 O3 , TiO2 , SiO2 ), carbides
(SiC, WC), diamond, solid lubricates (PTFE, graphite, MoS2 ), carbon
nano-tubes and even liquid-containing microcapsules. In general,
the presence of the particles (which are of a different phase) in
a co-deposited film enhances a variety of properties: microhardness, yield strength, tensile strength, wear and corrosion resistance,
self-lubrication, high temperature inertness and chemical and biological compatibility [1–12]. Metal-based composites have been
produced by other techniques, such as high velocity oxygen fuel
(HVOF), thermal spraying, plasma spraying, hot isostatic pressing, a
combination of physical and chemical vapor deposition and electroco-deposition. Because it is a technique conducted at a normal
pressure and ambient temperature, electrodeposition is considered
to be one of the most important techniques for producing composites [5]. In addition, electrodeposition is a low cost technique, has a
high deposition rate and leads to a homogenous distribution of the
particles. With the increasing availability of nanoparticles, there is
a growing interest in the electrolytic and electroless co-deposition
of nanoparticles [9,11–14].
Nickel which possesses high tensile strength and good toughness and corrosion resistance; it is a popular choice as a matrix
material because it can disperse both hard and soft reinforcements
to improve its wear and anti-friction resistance. Nickel composites
are noteworthy as an alternative to hard Cr because their enhanced
resistance to wear and high-temperature oxidation. Nickel plating
is widely adopted in automobile industries [14,15]. In addition, silicon carbide nickel matrix coatings have been extensively studied
and have gained widespread use for the protection against friction
inside parts of cylinders in the automotive industry. Considerable
research has been focused on the impact of the electrodeposition
parameters on the electrolytic co-deposition process of SiC with
Ni, as well as on the properties of the composite coatings. There
are various electrodeposition parameters: electrolysis conditions
(composition of the electrolytic bath, presence of additives and
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). In general, it has been observed that the amount of embedded SiC particles increases with both increasing concentration of
suspended SiC particles and increasing concentration of additives’
presence in the electrolyte [2,3,11,16–22].
The major challenges of the co-deposition of nanoparticles
seem to be the co-deposition of a sufficient number of nanoparticles, and avoiding the agglomeration of particles suspended
in the plating solutions. In the conventional methods of electrodeposition, one problem is that high nanoparticle agglomeration
occurs easily due to high surface activity of the nanoparticles
[9,24]. Homogeneity of the composite coating was promoted by
decreasing the ionic concentration of the electrolyte solution and
the using a specific ultrasonic energy treatment. The addition of
Table 1
Bath compositions and electrodeposition conditions for nano SiC-reinforced MMC
production.
Nickel sulfate (Ni2 SO4 ·6H2 O) (g/l)
Nickel chloride (NiCl2 ·6H2 O) (g/l)
Boric acid (H3 BO3 ) (g/l)
Sodyumdodecyl sulfate (g/l)
Cetyltrimethylammonium bromide (CTAB) (mg/l)
Silicon carbide (SiC) (g/l) (0.1–1 ␮m)
pH
Temperature (◦ C)
Current density (A/dm2 )
Stirring speed (rpm)
Plating time (h)
300
50
40
0.2
0, 100, 200, 300, 400
20
4
45
3
250, 650
2
metal cationic accelerants and organic surfactants in an electrolytic
bath improved the amount and the distribution of co-deposited
particles [14,23–25].
In this study, attempts have been made to develop composite coatings based on a Ni-SiC system using a SiC particle size
between 100 and 1000 nm. This work focuses on the effects of
stirring speed and surfactant content on the morphology of the
composite coatings, and the hardness and wear resistance of the
Ni SiC coatings is evaluated. Because the co-deposited SiC content
is a crucial factor for the coating properties, special attention was
paid to evaluate high volumes of SiC (e.g. 12 vol.%) in the coating
layer.
2. Experimental procedure
The processed plating electrolyte was a nickel sulfate bath for
which the composition, and the range of experimental process
parameters are shown in Table 1. The particle size range of the SiC
particles used in the experiment is 100–1000 nm. A nickel plate of
30 mm 35 mm was used as the anode, and a stationary iron substrate was used as the cathode. The bath was stirred by a magnetic
stirrer with a stirring speed between 250 and 650 rpm and was
heated to 45 ◦ C. the bath pH was fixed at 4. Prior to deposition,
the zeta potential of ceramic particles was measured with Malvern
Zetasizer Nano Series Nano-ZS model instrument.
After co-deposition, a scanning electron microscope (JEOL JSM 6060 LV) was used to observe the surface and the crosssection microstructures of the deposits. XRD analysis was carried
out at a speed of 1◦ /min in the 2 range of 10–100◦ with a Rigaku
D/MAX/2200/PC model device. The hardness of the coatings was
measured using a Vicker’s microhardness indenter (Leica VMHT)
with a load of 50 g. The reciprocating tribological behaviors of the
electrodeposited SiC nano composite coatings, sliding against M50
steel ball (Ø 10 mm) were examined on a CSM Instruments Tribometer designed according to DIN 50 324 and ASTM G 99-95a
in a ball-on-disk configuration. The wear tests were performed at
a constant applied load of 1 N with a sliding speed of 50 mm/s.
The morphologies of the wear traces were observed using SEMEDS analysis. The particle volume fractions were calculated directly
from the 6060 LV SEM image analysis program, which based on
phase area method.
3. Results and discussion
3.1. Effect of stirring speed on distribution of co-deposited SiC
particles
In this study, a number of stirring speeds were investigated
to optimize the deposition parameters. Here, two different stirring speeds are presented to show the effect of stirring speed on
the course of electrodeposition at a constant surfactant concentration of 300 mg/l. Hence, one low speed (250 rpm) and one high
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
55
Fig. 1. Cross-sectional SEM micrographs of the composite coatings showing the distribution of SiC particles that were deposited with the stirring speeds of (a) 250 rpm or
(b) 650 rpm.
speed (650 rpm) were chosen. Fig. 1 shows cross-sectional SEM
micrographs of the coatings deposited at these stirring speeds. The
volume percent of co-deposited SiC is found to be 2.54 vol.% at the
stirring speed of 250 rpm. Increasing the stirring speed to 650 rpm
led to an increase in the co-deposited SiC up to 10.05 vol. %. Because
the 650 rpm stirring speed resulted in producing segregation-free
and high SiC volume percentage co-deposited layers, the stirring
speed for all coatings was fixed at this level. Higher stirring speeds
were also studied in this work. Increasing stirring speed beyond
650 rpm resulted in obtaining layers with less SiC and an inhomogeneous distribution of SiC. Therefore, the results for higher stirring
speeds were not reported in this work.
3.2. Effect of surfactant (CTAB) concentration on the structure
Fig. 2 shows cross-sectional SEM micrographs of coatings
deposited with different surfactant (CTAB) concentrations in
the electrolyte. Increasing the CTAB concentration resulted in
increasing the volume percentage of co-deposited SiC particles. Introducing CTAB into the electrolyte not only increased
particle volume in the deposited layer but also resulted in a nonagglomerated dispersion of particles.
As seen in Fig. 3a, increasing the concentration of the surfactant
increased the zeta potential of the SiC particles. It has been reported
by Filiâtre et al. [26] that the zeta potential of SiC particles would
be increased by adsorption of the cationic surfactant CTAB. The
positive zeta potential offers an extra adhesion force between the
inert particles and the cathode and results in increasing the amount
of the embedded SiC particles. It is shown that the volume fraction of the co-deposited SiC can be increased up to approximately
11.5 vol.% by increasing the concentration of the surfactant CTAB
(Fig. 3b). As stated by Lee et al. [27], the adsorption of nickel ions
and protons in the plating bath on the SiC particle surface occurs,
which forms an ionic cloud around the SiC particles. The size of the
ionic cloud affects the electrophoretic behavior: the electrophoresis rate is higher with the increased adsorption of nickel ions and
protons.
The XRD analysis is a suitable method to determine the crystalline phase and presence of SiC nano-particles in the matrix of
Ni SiC MMCs. Fig. 4 shows the XRD patterns of pure nickel and
Ni SiC nano composites. The intensity of the (2 0 0) diffraction peak
of nickel in the nano composite coating is lower, and the peak width
is broader than that of the nickel coating (see Fig. 4b), as seen in
previous literature [12,28]. This is attributed to the decrease in the
grain size of the Ni SiC nano composite coating by the addition of
SiC nano-particles to the plating bath. SiC nano-particulates provide
more nucleation sites and hence retard the crystal growth; subsequently, the corresponding nickel matrix in the composite coating
has a smaller crystal size [12]. This may be correlated with [100]
texture, which is associated with deposits with minimum hardness and maximum ductility [3,28]. The Ni–SiC nano-composite
has exhibited increased (1 1 1), (2 2 2) and (3 1 1) diffraction lines
with an attenuation of the (2 0 0) line.
3.3. Effect of plating parameters on microhardness
The hardnesses of the unreinforced Ni coating and nano-SiC codeposited Ni-MMCs are shown in Fig. 5. It can be inferred that the
hardnesses of the co-deposited layer is highly correlated with the
volume fraction of SiC. According to Fig. 5, the enhanced hardness
results from the increased SiC content in the deposit. The microhardness of the Ni SiC composite coating containing 10.05 vol.%
SiC is 571 Hv while the microhardness of a pure nickel coating is
only 280 Hv.
It is interesting to note that, as Hou et al. [11] explained, the addition of surfactant in the electrolyte indeed significantly enhances
the hardness of the Ni deposit. Although the reason is not clear,
it is very encouraging to observe that this surfactant containing
Ni deposits possess higher hardness than pure Ni through this
electro-co-deposition process, because the enhanced wear property is expected from the hardened matrix [11]. Thus, the present
study would focus on the wear performance of well-dispersed SiC
content in the co-deposition layer.
3.4. Lattice distortion of nickel matrix
Lattice distortion of the Ni matrix for an unreinforced alloy
and for the Ni SiC, composite matrix was calculated with basic
reflections from the crystal planes. The lattice constants for unreinforced Ni coatings were also calculated and then compared with
the co-deposited Ni matrix reflections. The deviation in the lattice constants of the unreinforced Ni coatings and the composite
matrices were assessed as lattice distortion. For an f.c.c. system,
the lattice constant can be calculated as follows [29]:
a2 = (
2
4 sin2 )(h2 + k2 + l2 )
(1)
The deviation in the lattice constant, depending on the particle
content and deposition parameters is calculated with the equation,
Distortion% = 100[
a0 − a1
]
a0
(2)
where a0 is the lattice constant of the unreinforced Ni coating and
a1 is the lattice constant of the Ni matrix for composites produced
under different conditions.
56
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
30
Volume percentage of SiC (vol. %)
Fig. 2. SEM micrographs of the distribution of SiC particles in the cross-section of composite coatings that were prepared with the following concentrations of CTAB; (a)
0 mg/l, (b) 100 mg/l, (c) 200 mg/l, (d) 300 mg/l or (e) 400 mg/l.
(a)
Zeta potential (mV)
20
10
0
-10
-20
-30
0
100
200
300
400
Concentration of surfactant (CTAB) (mg/l)
12
(b)
10
8
6
4
2
0
0
100
200
300
400
CTAB concentration (mg/l)
Fig. 3. (a) CTAB concentration–zeta potential relationship, (b) the volume percentage of co-deposited SiC particles for various concentrations of CTAB.
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
57
0,5
% Distortion
0,4
0,3
0,2
0,1
0,0
(111)
(200)
(311)
-0,1
0
100
200
300
400
500
CTAB concentration (mg/l)
Fig. 6. Effect of the CTAB concentration in the electrolyte on the lattice distortion
of the Ni matrix.
Fig. 4. Effect of various concentrations of particles in the electrolyte on XRD patterns.
The effect of the surfactant (CTAB) concentration on the lattice
distortion was calculated and is demonstrated in Fig. 6. Increasing
the CTAB initially showed a positive lattice distortion behavior up
to a 300 mg/l concentration. Beyond this surfactant concentration,
the lattice distortion values showed a sharp decrease in a negative
way.
3.5. Wear and friction
3.5.1. Effect of surfactant on wear and friction properties
The effect of the surfactant concentration in the electrolyte was
studied at a constant particle concentration of 20 g/l. It is noted
that free cationic surfactant CTAB embedded in the nickel matrix
resulted in a hardened matrix; therefore, there was a decrease in
the wear track as shown in Fig. 7. The wear tracks in Fig. 7 are as
follows: 7a (0 mg/l CTAB in the electrolyte), 7b (100 mg/l CTAB in
the electrolyte), 7c (200 mg/l CTAB in the electrolyte), 7d (300 mg/l
CTAB in the electrolyte) and 7e (400 mg/l CTAB in the electrolyte).
Fig. 7 shows the morphology of the worn surfaces of the composite coatings deposited at various surfactant (CTAB) concentrations.
Plastic deformation of the coatings was observed in all cases. However, the degree of plastic deformation was reduced for the coatings
600
10,05%
11,37%
Microhardness (HV)
500
8,47%
400
1,26%
1,67%
0
100
300
200
100
0
Ni
200
300
400
CTAB concentration (mg/l)
Fig. 5. Effect of the CTAB concentration in the electrolyte on microhardness. The
percentage values are presented to show particle volume in the deposited layer.
deposited at high CTAB concentrations, except for the coatings that
were produced with 400 mg/l CTAB.
The width of the wear track produced on the coated surfaces was
used as a qualitative measure to compare the wear resistance of the
composite coatings. Increasing the CTAB concentration in the electrolyte resulted in smaller and shallower wear tracks. The width of
the wear track was decreased up to 300 mg/l CTAB addition. Beyond
this value, the surface damage was increased, and the wear tracks
showed again larger widths and depths. It was found that the width
of the wear track of the coatings deposited with 300 mg/l CTAB in
the electrolyte was smaller than the widths of the wear tracks of
the coatings deposited with all other CTAB concentrations. As mentioned earlier, higher contents of SiC particles in the coatings were
detected for the coatings deposited with the increased amount of
CTAB. A large volume fraction of reinforced particles in the metal
matrix has the effect of reducing the distance between the particles in the matrix and thus can make the coating harder and
more resistant to plastic deformation. Consequently, it becomes
more difficult for the counter ball to remove the reinforced particles from the coating, and the lack of plastic deformation increases
the wear resistance of the composite coating. It is possible that the
surfactant CTAB in the electrolyte would significantly decrease the
agglomeration of SiC particles; therefore, a high amount of embedded SiC particles can be achieved. Furthermore, the present results
show the SiC composition in the co-deposition layer is enhanced
by the well-dispersed and relatively smaller SiC particles in the
electrolyte, which have a higher chance of being embedded in the
cathode. However, when the CTAB concentration is increased to
400 mg/l in the electrolyte, the wear loss is increased as shown in
Figs. 7e and 8a. The EDS compositional analysis on different areas
of the wear surface detected some amounts of iron, molybdenum
and chromium, which indicated that some debris was transferred
from the counter ball M50 to the surface. Microscopic examination
of the worn surface of the counter ball observed a round-shaped
wear scar with numerous thin and smooth scratches. This observation suggests that the counter ball experienced a polishing wear
due to the abrasive action of hard SiC particles of the composite
coatings. The hardness of the counter ball is lower than that of SiC
particles, and the SiC particles could plough into the surface of the
counter ball. Increasing the amount of SiC in the deposited layer
resulted in increasing the elemental transfer from the steel ball. The
highest amount of iron, molybdenum and chromium was detected
on the worn surface of the sample that was produced with 400 mg/l
CTAB. This result was attributed to poor interface bonding.
Fig. 8 shows the variation in wear rate and the coefficient of
friction of the Ni–SiC nano composite coatings with the varying
58
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
Fig. 7. SEM morphology of the wear tracks of composite coatings prepared with the following concentrations of CTAB: (a) 0 mg/l CTAB (1.26 vol.% SiC), (b) 100 mg/l CTAB
(1.67 vol.% SiC), (c) 200 mg/l CTAB (8.47 vol.% SiC), (d) 300 mg/l CTAB (10.05 vol.% SiC) and (e) 400 mg/l CTAB (11.37 vol.% SiC).
(b)
20
1,26%
16
Ni
1,67%
O
12
8,47%
11,37%
8
10,05%
4
0
Ni
0
100
200
CTAB (mg/l)
300
400
Coefficient of Friction (η')
Wear Rate (x10-4 mm3/Nm)
(a)
1,6
1,4
1,2
1,0
0,8
0,6
0,4
0,2
0,0
-100
0
100
200
300
400
500
CTAB (mg/l)
Fig. 8. Effect of the concentration of CTAB on (a) wear rate and (b) friction coefficient of Ni–SiC composite coatings.
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
59
Fig. 9. SEM morphology of the wear tracks of composite coatings prepared with the following concentrations of CTAB: (a) 0 mg/l CTAB (1.26 vol.% SiC), (b) 100 mg/l CTAB
(1.67 vol.% SiC), (c) 200 mg/l CTAB (8.47 vol.% SiC), (d) 300 mg/l CTAB (10.05 vol.% SiC) and (e) 400 mg/l CTAB (11.37 vol.% SiC).
concentration of CTAB. It is seen in Fig. 8a that the Ni–10.05 %
SiC nano composite coating has a minimum wear rate, which is
in accordance with Archard’s law [30] (compare Figs. 5 and 8).
As previously stated, the microhardness of the nanocomposite
coatings increases with an increasing percentage of the nano-SiC,
up to 300 mg/l CTAB concentration.
The increase in the microhardness and the decrease in the wear
rate of the Ni–SiC nanocomposite coatings (as compared to the
Ni coating) can be explained because the SiC nanoparticles codeposited in the Ni matrix can restrain the growth of the Ni grains
and the plastic deformation of the matrix under a loading by way
of grain-refining and dispersive strengthening effects. The grainrefining and dispersive strengthening effects become stronger with
increasing nano-SiC content; thus, the microhardness and wear
resistance of the Ni-SiC composite coatings increase with increasing nano-SiC content [30]. However, there is no obvious relation
between the coefficient of friction and CTAB, as shown in Fig 8b.
It can be concluded that introducing higher amounts of SiC by
increasing the addition of CTAB in the electrolyte does not affect
the coefficient of friction. In general, a large tendency for plastic
deformation of asperity junctions results in a higher and unstable
friction coefficient.
SEM pictures of wear tracks on electrodeposited Ni–SiC coatings
are shown in Fig. 9 at high magnification. Cracking and spalling
can be seen on the worn surface of the Ni coating produced without CTAB addition to the electrolyte (Fig. 9a). The presence of the
cracking and spalling causes much wear loss. The results suggest
that the load-bearing capacity, and the wear resistance of the CTABfree deposited Ni film are rather weak. Addition of 100 mg/l CTAB
produced a similar worn surface. The composite that was produced
with 100 mg/l CTAB in the electrolyte showed a similar worn surface, but a smaller amount of plastic deformation associated with
some small cracks (Fig. 9b). In the morphology of the worn surfaces
of the samples produced with 200 and 300 mg/l CTAB addition, the
worn areas are quite smooth. The appearance of all wear tracks of
the coatings produced with 200 and 300 mg/l CTAB addition was
60
F. Kılıc¸ et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 419 (2013) 53–60
similar and not dependent on the content of SiC particles in the
composite coatings (Fig. 9c and d).
The presence of wear debris will cause larger wearing losses.
The anti-wear performance of the composite produced with the
400 mg/l CTAB addition to the electrolyte is poorer than that of
the composite coating produced with the 300 mg/l CTAB addition.
Fig. 9e is the wear track morphology of the composite coating produced with the 400 mg/l CTAB concentration. Abrasive grooves can
be found in the direction parallel to the sliding, and a considerable
amount of debris can also be found in the wear track. It is possible
that the formation of the debris is due to weak bonding between
the agglomerated particles and the nickel matrix. Under the electrodeposition conditions, the deposited nickel likely cannot bind
the agglomerated particles tightly, and this may increase the brittleness of the coatings. As a result, the wearing counter body can
easily damage the surface of the coatings, and consequently, the
wear weight losses increase.
The experimental result showed that the incorporation of SiC
particles in the particle range of 100–1000 nm in the matrix can
largely improve the tribological performance of the co-deposited
Ni SiC composite coatings. It is well known that the hardness and
other mechanical properties of metal matrix composites depend
in general on the amount and size of the dispersed phase (apart
from the mechanical characteristics of the matrix). The amount
and size of particles define two kinds of reinforcing mechanisms
in metal matrix composite materials: dispersion- strengthening
and particle-strengthening. Because the particle range in this
study is between 100 and 1000 nm, the dispersion-strengthening
mechanism can play the main role through a dislocation–particle
interaction or Orowan hardening mechanism. Strengthening is
achieved because particles restrain deformation of the matrix by
a mechanical constraint.
4. Conclusions
1. SiC nanoparticle-reinforced (100–1000 nm) Ni metal matrix
composites were successfully produced by D.C. electroplating,
up to 11.37 vol.% particle co-deposition.
2. Increasing the surfactant (CTAB) content resulted in an increase
of SiC vol.% within the Ni matrix and a segregation-free dispersion of nanoparticle deposition.
3. The hardness values of the nano SiC-reinforced electrodeposited
coatings were as high as 571 Hv because of unique dispersion
effects.
4. It was observed that the volume percentage of SiC in coatings
increases with increasing surfactant (CTAB) concentration.
5. Increasing the surfactant (CTAB) concentration resulted in the
lattice distortion of the Ni matrix.
6. Wear resistance of the coatings was increased with increasing
surfactant (CTAB) content up to 300 mg/l in the electrolyte, but
beyond this concentration, the wear resistance decreased.
7. The co-deposited Ni–SiC nanocomposite coatings show higher
friction coefficients and better wear resistance compared to
the as-deposited Ni film, which can be attributed to the incorporation of nano-sized SiC particles in the deposit. These
nanoparticles greatly increase the hardness of the composite
coating through grain refinement strengthening and dispersionstrengthening mechanisms.
Acknowledgments
This work is supported by the Scientific and Technical Research
council of Turkey (TUBITAK) under contract number 106M253. The
authors thank the TUBITAK MAG workers for their financial support.
References
[1] X.B. Shi, W.H. Dou, D.S. Yun, J. Bin, T. Wei-yi, Electrodepositing nickel silica
nano-composites coatings, Electrochem. Commun. 7 (2005) 572–575.
[2] M.D. Ger, Electrochemical deposition of nickel/SiC composites in the presence
of surfactants, Mater. Phys. Chem. 87 (2004) 67-–74.
[3] P. Gyftou, E.A. Pavlatou, N. Spyrellis, Effect of pulse electrodeposition parameters on the properties of Ni/nano-SiC composites, Appl. Surf. Sci. 254 (2008)
5910–5916.
[4] M. Surender, R. Balasubramaniam, B. Basu, Electrochemical behavior of electrodeposited Ni–WC composite coatings, Surf. Coat. Technol. 187 (2004) 93–97.
[5] L. Shi, C. Sun, W. Liu, Electrodeposited nickel–cobalt composite coating containing MoS2 , Appl. Surf. Sci. 254 (2008) 6880–6885.
[6] L. Shi, C.F. Sun, F. Zhou, W.M. Liu, Electrodeposited nickel–cobalt composite
coating containing nano-sized Si3 N4 , Mater. Sci. Eng. A 397 (2005) 190–194.
´
˙ W. Szymanski,
´
E. Rudnik, J. Koza, L. Błaz,
Electrodeposition and
[7] L. Burzynska,
heat treatment of nickel/silicon carbide composites, Surf. Coat. Technol. 202
(2008) 2545–2556.
[8] Y. Yao, S. Yao, L. Zhang, H. Wang, Electrodeposition and mechanical and corrosion resistance properties of Ni–W/SiC nanocomposite coatings, Mater. Lett.
61 (2007) 67–70.
[9] I. Garcia, J. Fransaer, J.-P. Celis, Electrodeposition and sliding wear resistance of
nickel composite coatings containing micron and submicron SiC particles, Surf.
Coat. Technol. 148 (2001) 171–178.
[10] E. Pompei, L. Magagnin, N. Lecis, P.L. Cavallotti, Electrodeposition of nickel–BN
composite coatings, Electrochem. Acta 54 (2009) 2571–2574.
[11] K.H. Hou, M.D. Ger, L.M. Wang, S.T. Ke, The wear behaviour of electrocodeposited Ni–SiC composites, Wear 253 (2002) 994–1003.
[12] M.R. Vaezi, S.K. Sadrnezhaad, L. Nikzad, Electrodeposition of Ni–SiC nanocomposite coatings and evaluation of wear and corrosion resistance and
electroplating characteristics, Colloids. Surf. A: Physicochem. Eng. Aspects 315
(2008) 176–182.
[13] M. Srivastava, V.K.W. Grips, A. Jain b, K.S. Rajam, Influence of SiC particle size
on the structure and tribological properties of Ni–Co composites, Surf. Coat.
Technol. 202 (2007) 310–318.
[14] C.S. Ramesh, S.K. Seshadri, Tribological characteristics of nickel based composite coatings, Wear 255 (2003) 893–902.
[15] L. Orlovskaja, N. Periene, M. Kurtinaitiene, G. Bikulcius, Electocomposites with
SiC content modulated in layers, Surf. Coat. Technol. 105 (1998) 8–12.
[16] Q. Feng, T. Li, H. Yue, K. Qi, F. Bai, J. Jin, Preparation and characterization of
nickel nano-Al2O3 composite coatings by sediment co-deposition, Appl. Surf.
Sci. 254 (2008) 2262–2268.
[17] S.L. Kuo, Y.C. Chenb, M.D. Ger, W.H. Hwu, Nano-particles dispersion effect on
Ni/ Al2O3 composite coatings, Mater. Phys. Chem. 86 (2004) 5–10.
[18] S.C. Wang, W.C.J. Wei, Kinetics of electroplating process of nano-sized ceramic
particle/Ni composite, Mater. Phys. Chem. 78 (2003) 574–580.
[19] C.T.J. Low, R.G.A. Wills, F.C. Walsh, Electrodeposition of composite coatings
containing nanoparticles in a metal deposit, Surf. Coat. Technol. 201 (2006)
371–383.
[20] L. Orlovskaja, N. Periene, M. Kurtinaitiene, S. Surviliene, Ni–SiC composite
plated under a modulated current, Surf. Coat. Technol. 111 (1999) 234–239.
[21] A.F. Zimmerman, G. Palumbo, K.T. Aust, U. Erb, Mechanical properties of nickel
silicon carbide nanocomposites, Mater. Sci. Eng. A 328 (2002) 137–146.
[22] I. Garcia, A. Conde, G. Langelaan, J. Fransaer, J.P. Celis, Improved corrosion
resistance through microstructural modifications induced by codepositing SiCparticles with electrolytic nickel, Corr. Sci. 45 (2003) 1173–1189.
[23] Z.X. Niu, F.H. Cao, W. Wang, Z. Zhang, J.Q. Zhang, C.N. Cao, Electrodeposition of
Ni-SiC nanocomposite film, Trans. Nonferrous Met. Soc. China 17 (2007) 9–15.
[24] L. Chen, L. Wang, Z. Zeng, J. Zhang, Effect of surfactant on the electrodeposition
and wear resistance of Ni– Al2 O3 composite coatings, Mater. Sci. Eng. A 434
(2006) 319–325.
[25] T. Lampke, B. Wielage, D. Dietrich, A. Leopold, Details of crystalline growth in
co-deposited electroplated nickel films with hard (nano)particles, Appl. Surf.
Sci. 253 (2006) 2399–2408.
´
Electrodeposition
[26] C. Filiâtre, C. Pignolet, A. Foissy, M. Zembala, P. Warszynski,
of particles at nickel electrode surface in a laminar flow cell, Colloids Surf. A:
Physicochem. Eng. Aspects 222 (2003) 55–63.
[27] H.K. Lee, H.Y. Lee, J.M. Jeon, Codeposition of micro- and nano-sized SiC particles
in the nickel matrix composite coatings obtained by electroplating, Surf. Coat.
Technol. 201 (2007) 4711–4717.
[28] M. Srivastava, V.K. William Grips, K.S. Rajam, Electrochemical deposition and
tribological behaviour of Ni and Ni–Co metal matrix composites with SiC nanoparticles, Appl. Surf. Sci. 253 (2007) 3814–3824.
[29] I. Misbah-Ul, K.A. Hashmi, M.U. Rana, T. Abbas, Measurement of exchange interaction in Ti-substituted Ni-ferrites, Solid State Commun. 121 (2002) 51–54.
[30] L. Shi, C. Sun, P. Gao, F. Zhou, W. Liu, Mechanical properties and wear and corrosion resistance of electrodeposited Ni–Co/SiC nanocomposite coating, App.
Surf. Sci. 252 (2006) 3591–3599.