You are on page 1of 8

Journal of Alloys and Compounds 635 (2015) 150–157

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


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

Composition, characteristics and tribological behavior of Cr, Co–Cr


and Co–Cr/TiO2 nano-composite coatings electrodeposited
from trivalent chromium based baths
S. Mahdavi, S.R. Allahkaram ⇑
School of Metallurgy and Materials Engineering, University College of Engineering, University of Tehran, Tehran, Iran

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

Article history: Cr, Co–Cr and Co–Cr/TiO2 nano-composite coatings were electrodeposited from Cr(III) based electrolytes
Received 14 January 2015 onto 316L SS substrates. The effects of TiO2 nano-particles concentration on co-deposition of these par-
Accepted 15 February 2015 ticles along with Cr content and microhardness of the coatings were investigated. Morphology and struc-
Available online 21 February 2015
ture of the Cr, Co–Cr and Co–Cr/TiO2 coatings besides their tribological behavior were studied, and the
results were compared to the substrate. The results showed that co-deposition of TiO2 nano-particles
Keywords: and chromium content of the coatings increased with enhancing particles concentration in the bath.
Co–Cr alloy
Microhardness of Co–Cr and Co–Cr/TiO2 coatings were improved by increasing their Cr and TiO2 content,
Electrodeposition
Nano-composite
respectively. However, the presence of more than 3 wt.% TiO2 in the composite films had a negligible
Trivalent chromium effect on microhardness of these samples. All the coatings had nodular morphology and contained micro-
Titania cracks. Nodules sizes and the number of microcracks in the alloy and composite coatings were lower than
Tribological behavior the Cr film. The structure of the Co–Cr and Co–Cr/TiO2 coatings were crystalline, while it was amorphous
for Cr film. Wear results revealed that the Co–29.1Cr/3.0TiO2 coating had the lowest wear loss and coef-
ficient of friction between all the samples. The wear resistance of the Co–29.3Cr coating was better than
the substrate, while Cr film had the worst wear resistance. According to SEM micrographs from the worn
surface of the samples, abrasion was the main wear mechanism in 316L SS substrate, as well as in the Co–
29.3Cr and Co–29.1Cr/3.0TiO2 coatings. However, limited adhesion also occurred on the worn surface of
the substrate. Detachment of Cr from the wearing surface during the sliding resulted in higher volume
losses in this sample.
Ó 2015 Elsevier B.V. All rights reserved.

1. Introduction and corrosion resistant coatings with high hardness and good
appearance [4,10–16]. Conventional chromium coatings were
Engineering applications of metals can be limited by surface deposited from hexavalent chromic acid solutions for many years
failures due to wear, corrosion, fatigue, etc. Therefore, surface [14–20]. However, Cr(VI) plating should be replaced by alternative
modification can increase service life and reliability of metallic processes, because hexavalent chromium is carcinogenic and toxic
components [1]. Different surface coating approaches such as plas- [10–21]. Unfortunately, a few coatings could be an alternative for
ma spraying, sol–gel, high velocity oxygen fuel (HVOF), chemical the conventional hard chromium films [16,18]. In recent years;
vapor deposition (CVD), physical vapor deposition (PVD), laser electrodeposition of Cr from trivalent chromium electrolytes is
cladding, ion implantation, and electrodeposition have been used considered to be a convenient and low cost method to replace con-
in order to improve surface properties [2–5]. Among these meth- ventional Cr(VI) plating [10,12–24]. However, Cr(III) based coatings
ods; electrodeposition is a simple and economic technique for pre- have lower corrosion and wear resistance than those obtained
cision coating of different shaped substrates with pure metal, alloy from Cr(VI) electrolytes [21,25]. Properties of these coatings can
or composite films [4–9]. be improved by adding iron family metal ions (i.e. iron, nickel,
Electrodeposition of chromium is widely used in aerospace, and Co) to the Cr(III) bath and electrodeposition of M-Cr alloys
automotive, and other engineering industries to produce wear [4,14,16–18,24–26].
Co–Cr alloys are used for medical prosthetic implant devices,
components in the power generation, marine, aerospace, and oil
⇑ Corresponding author. Tel.: +98 2161114094. and gas industries due to their good wear, corrosion and fatigue
E-mail address: akaram@ut.ac.ir (S.R. Allahkaram). resistance along with biocompatibility [3,4,23,27–32]. Optimum

http://dx.doi.org/10.1016/j.jallcom.2015.02.119
0925-8388/Ó 2015 Elsevier B.V. All rights reserved.
S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157 151

mechanical properties and corrosion resistance of these alloys are A double electrode cell was used for electroplating experiments. 316L stainless
steel plates with a thickness of 2.5 mm and exposed area of 9 cm2 (3  3 cm) were
usually obtained in the presence of 10–30% chromium [23]. Czako-
used as the substrate. A similar dimension of a Ti/IrO2 dimensionally stable anode
nagy et al. [33] have used DC current for electrodeposition of Co–Cr was used as the anode that was vertically maintained at 3 cm from the cathode. Pri-
alloys from Cr(III) based electrolytes. They reported that chromium or to electrodeposition, the substrates were abraded with different grades of emery
content of the alloys was diminished by increasing of temperature papers from 120 to 2000 grit which were then ultrasonically cleaned in acetone for
or decreasing of current density. Efimov and Chernykh [26] have 10 min at room temperature, washed in distilled water and degreased in an alkaline
solution (NaOH, Na2CO3, Na2PO412H2O) at 70 °C for 15 min. Activation of the 316L
studied the effect of electrolyte composition, pH of the solution,
SS cathodes was done in a solution containing combination of HCl and CoCl26H2O
and the electrolysis conditions on the Co–Cr alloys composition at room temperature, according to ASTM B254. With this technique, the oxide layer
and the coatings quality. They reached to maximum chromium on the surface of 316L SS is removed, and a thin cobalt layer is deposited onto the
content of 15% within the alloy, and found that Co–Cr alloy coat- surface. According to previous researches, this layer can improve adhesion and
properties of chromium contained electrodeposits on steel substrates [20]. The sub-
ings had better corrosion resistance than Cr film. Surviliene et al.
strates were transferred to the electroplating bath after activation, immediately.
[18] have produced Co–Cr alloy coatings containing different per- Hardness measurements were carried out on an AMSLER D-6700 Vicker’s
centages of Co by adding various concentrations of CoCl2 to the microhardness testing machine, using an applied load of 1 N for 15 s. The mean val-
Cr(III) formate–urea baths and examined the chemical composition ues of at least five measurements conducted at various locations on each sample
of the top layers by means of XPS. According to Ohgai et al. [19] were considered.
Dry sliding ball-on-disc wear tests on the substrate and coated samples were
results, the electrodeposition process of Co–Cr was categorized to
carried out in a laboratory atmosphere with 30–40% relative humidity and at room
normal co-deposition type, and Co-8.4% alloys formed substitu- temperature. An alumina ball with a diameter of 6 mm was used as the counterface
tional solid solution phase with fine crystals. Corrosion properties (Supplied by Ardakan Industrial Ceramics, which had hardness and maximum sur-
of Co–(19–35%)Cr alloys, deposited from deep eutectic solvents face roughness of 2000 Hv and 0.02 lm Ra, respectively). Wear tests were per-
formed under the applied load of 8 N, with the sliding distance of 300 m, at a
containing CoCl2, and CrCl3 without any Cr complexing agent, were
sliding speed of 0.1 m s 1 and a wear track diameter of 1.6 mm. Both ball and disc
investigated by Saravanan and Mohan [23]. They found that Co– samples were cleaned in an ultrasonic acetone bath for 2 min and dried with com-
35%Cr was more corrosion resistant than Co–19%Cr alloy. pressed air before and after testing. The lateral forces were also recorded during the
Although some research on electrodeposition of Co–Cr alloys wear tests and friction coefficients of the samples were calculated.
from Cr(III) baths as well as characterization and investigation of The electrodeposited and worn surface morphologies of the coated samples,
along with the amount of co-deposited TiO2 nano-particles within the Co–Cr matrix
their magnetic and corrosion properties have been conducted,
and worn surface composition were examined using a scanning electron micro-
there are limited studies about electrodeposition of Co–Cr alloys scope equipped with energy dispersive X-ray spectrometer (EDS).
from chromium sulfate based electrolytes and examination of their
tribological behavior. Moreover, most of the previous research
3. Results and discussion
have been focused on production of unreinforced Co–Cr alloys,
while co-deposition of fine particles into metallic electrodeposits
3.1. Bath TiO2 nano-particles concentration
as the second phase results in enhanced hardness, corrosion and
wear resistance of the metallic coatings [4–9,12,21,34–38]. TiO2
is a bioactive ceramic material [5,39], which can improve the coat-
3.1.1. Effect on coatings composition
ings tribological behavior and corrosion resistance. In the present
The effect of TiO2 nano-particles in the bath on the TiO2 and
study, DC electrodeposition technique was used to produce Co–
chromium content of the deposits is shown in Fig. 1. Volume frac-
Cr and Co–Cr/TiO2 nano-composite coatings. The aim of this study
tion of incorporated TiO2 nano-particles promotes from 0 to
is to investigate the morphology, microstructure, hardness, and tri-
3.5 wt.% (0 to 6.6 vol.%) with an increase of the bath TiO2 concen-
bological behavior of these coatings and compare the results with
tration from 0 to 30 g L 1, tending to attain a steady value at high
electrodeposited Cr film and the substrate.
TiO2 concentrations. The curve is similar to the Langmuir adsorp-
2. Experimental
tion isotherm, which is consistent with previous investigations
and Guglielmi’s model of adsorption [5,7,8,36,40]. According to
Cr, Co–Cr and Co–Cr/TiO2 coatings with average thickness of about 25 lm were this model; the first step is adsorption of metal ions onto the
produced by DC electrodeposition from chromium sulfate based electrolytes. The ceramic particles and loose physical adsorption of the particles
bath compositions and plating conditions are presented in Table 1. Analytical grade
onto the cathode. The second step is the strong adsorption and
chemicals and double distilled water were used for preparation of the electrolytes.
The pH value of the bath was adjusted by adding either 4 M H2SO4 or 1 M NaOH entrapment of the particles within the metal matrix. Increasing
solution. Prior to composite plating, 24 h stirring along with 45 min ultrasonic the concentration of TiO2 particles in the solution can in turn raise
treatment were used for better dispersion of the nano-particles within the electro- the amount of loosely adsorbed particles on the cathode surface.
plating bath.

Table 1
Bath composition and plating condition for electrodeposition of Cr, Cr–Co and Co–Cr/
TiO2 coatings.

Bath composition Cr coating Co–Cr and Co–Cr/TiO2


coatings
Cr2(SO4)36H2O 0.3 M 0.35 M
CoSO47H2O – 0.1 M
H3BO3 0.9 M 0.9 M
Na2SO4 0.6 M 0.6 M
C3H4O4 (Malonic acid) 0.5 M 0.6 M
C6H8O7 (Citric acid) – 0.4 M
1
NaC12H25SO4 (SDS) 0.2 g L 0.2 g L 1
1
TiO2 (average size of about 30 nm) – 0–30 g L
Deposition condition
2 2
Current density 140 mA cm 140, 150 mA cm
Temperature 30 ± 1 °C 30 ± 1 °C
pH 1.5 1
Stirring speed 250 rpm 250 rpm Fig. 1. Effect of bath TiO2 nano-particles concentration on weight percent of co-
Deposition time 60 min 30 min deposited particles and Cr content of the coatings, electrodepositing at
140 mA cm 2.
152 S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157

Hence, higher volume fractions of the nano-particles can be strong- 3.5TiO2 coatings, obtaining from the solutions having 20 and
ly adsorbed at the next step. 30 g L 1 nano-particles, confirms agglomeration of nano-particles
Fig. 1 also shows that chromium content of the coatings is in the second sample (Fig. 3). The balance between agglomeration
enhanced by increasing the bath TiO2 concentration. The chromi- of the nano-particles in the composite coatings containing more
um content of the Co–Cr coating is 23.7 wt.%, while it is than 3.0 wt.% TiO2 and increasing amount of Cr is almost constant
31.2 wt.% for Co–Cr/3.5TiO2 film. According to the previous studies, hardness in these films.
alloy coatings with higher chromium contents are obtained by Considering the obtained results, the Co–29.1Cr/3.0TiO2 coating
increasing cathodic current density, which is due to the variation is selected as the most appropriate composite film for further
in the rate constants of electrode reactions [14,22,24,33,41,42]. investigations. Morphology and tribological behavior of Cr and
Our results also show that, the chromium percentage of the Co– Co–29.3Cr coatings are also examined, and the results are com-
Cr alloy coatings rises from 23.7 to 29.3 wt.% with increasing the pared to the composite film. From now on, the selected alloy and
current density from 140 to 150 mA cm 2. Co-deposition of inert composite coatings are briefly called Co–Cr and Co–Cr/TiO2.
particles can cause surface blockage and reduce electroactive elec-
trode area, resulting in an increment of current density on the
3.2. Morphological and structural investigation
cathode surface [7,34,43]. Therefore, higher chromium content of
the composite coatings as compared to the alloy film can be
SEM micrographs from the surface of the Cr, Co–Cr and Co–Cr/
attributed to the reduced active electrode area and increased
TiO2 coatings are presented in Fig. 4. Deposits show nodular sur-
cathodic current densities.
face morphology with microcracks. Nodules size along with the
number and size of microcracks in Cr coating are greater than
3.1.2. Effect on coatings microhardness the alloy and the composite films. Similar results have also been
Hardness test is one of the relatively non-destructive, simple obtained by Survilienė et al. [18]. Formation of cracks is character-
and easy techniques to investigate changes in mechanical proper- istic of electrodeposited Cr coatings. Cracks are formed due to
ties by variation of the chemical composition of the alloys [44]. The hydrogen inclusion and incorporation of basic chromium com-
microhardness of the produced coatings is shown in Fig. 2. Hard- pounds, which develop internal stresses in the deposited films
ness value of the substrate is also presented for the purpose of and deteriorate the mechanical properties [16,22,24,41]. Co-depo-
comparison. It is clear that all the coatings have higher microhard- sition of Co with Cr can decrease both hydrogenation and the
ness as compared to the substrate. The hardness values for elec- incorporation of basic chromium compounds. Thus, internal stress-
trodeposited Co–23.7Cr and Co–29.3Cr are 562 and 596, es are reduced [16,24,41], resulting in lower cracks density of Co–
respectively. These hardness values are higher than that of the bulk Cr alloy coating.
and sprayed Co–Cr–Mo alloys [3,28,29,32,45]. It is also clear from Fig. 4 that adding nano-particles to the Co–
Microhardness of Co–Cr coatings lies between the hardness val- Cr coating has a negligible effect on the nodules size, while slightly
ues of Cr and pure nano-crystalline cobalt that was previously
deposited (about 470 Hv) [5]. The hardness of Cr is greater than
that of Co, thus harder coatings accompanied by higher Cr content.
A similar trend has also been observed by other researchers for Co–
Cr and Ni–Cr coatings [41,42,46].
The hardness of the coatings elevates with increasing their
nano-particles content to a certain value at 3 wt.%, after which
remains almost constant. Hardness improvement by incorporation
of TiO2 nano-particles is related to dispersion strengthening effect,
which inhibits plastic deformation of the metal matrix by prevent-
ing free motion of the dislocations [5–8,46]. Moreover, increasing
of Cr content of the coatings by adding TiO2 nano-particles could
be another reason of hardness increment. However, agglomeration
of the nano-particles in the presence of high concentrations of
them in the bath may weaken dispersion strengthening effect of
the nano-particles and decrease coatings hardness [9,31]. Compar-
ing SEM micrographs on cross-section of Co–Cr/3.0TiO2 and Co–Cr/

Fig. 3. SEM micrographs from cross-section of (a) Co–29.1Cr/3.0TiO2 and (b) Co–
Fig. 2. Microhardness values of different samples. 31.2Cr/3.5TiO2 composite coatings.
S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157 153

Fig. 5. XRD patterns of (a) Cr, (b) Co–Cr and (c) Co–Cr/TiO2 coatings.

may be a result of microcracks, which expose the substrate to the


X-ray.
The structure of the alloy and composite coatings is a solid solu-
tion of Cr in Co with hexagonal close-packed (hcp) lattice. Broad
picks are a result of ultra-fine crystallite sizes of these samples.
The structure of Co–Cr alloy coating is not changed by co-deposi-
tion of TiO2 particles.

3.3. Tribological behavior

3.3.1. Wear properties


Wear volume losses of 316L SS substrate, Cr, Co–Cr and Co–Cr/
TiO2 coatings are exhibited in Fig. 6. Clearly, the Co–Cr and Co–Cr/
TiO2 coatings have better wear resistance than the substrate. The
volume losses of the alloy and the nano-composite films are about
3 and 7.5 times lower than the 316L SS substrate, respectively.
According to the Archard theory, wear resistance is proportional
to the hardness of the material [5,47]. Given hardness values of
the samples (Fig. 2), better wear resistance of the Co–Cr and Co–
Cr/TiO2 coatings is due to their higher hardness as compared to
the 316L SS substrate.
Fig. 6 also shows that the wear behavior of Co–Cr alloy coating
is improved by incorporation of TiO2 nano-particles, which is con-
sistent with the results of other investigators [5–7,15,35,46].
Enhanced wear resistance of nano-composite coating as compared
to the alloy film can be a result of its higher hardness along with
the dispersion strengthening effect of TiO2 nano-particles, which
prevents grains movement and grain boundaries migration during
Fig. 4. SEM micrographs from the surface of the as-deposited (a) Cr, (b) Co–Cr and the wear test [5,48].
(c) Co–Cr/TiO2 coatings. The wear volume loss of the Cr film is about 13% higher than the
substrate, although Cr coating is significantly harder than the 316L
decreases the number of microcracks. Hydrogenation and residual
stresses of the coatings are not affected by co-deposition of parti-
cles [12]. Therefore, the lower cracks number in the Co–Cr/TiO2
nano-composite coating as compared to the Co–Cr film can be
attributed to the increased toughness of the composite coating
by incorporation of particles, as also reported by other researchers
[15,37,38].
XRD patterns of the coatings are presented in Fig. 5. These pat-
terns reveal that the as-deposited Cr film has an amorphous struc-
ture, whereas Co–Cr and Co–Cr/TiO2 coatings have a crystalline
structure. Formation of amorphous Cr electrodeposits from Cr(III)
baths has also been reported by other researchers [16,20,21].
Application of these amorphous coatings in the aspect of anti-wear
is limited due to the brittleness of amorphous materials, which can
be improved by crystallization [21]. The small sharp peak in the
XRD pattern of Cr film at 74° is related to the 316L SS substrate. Fig. 6. Wear volume losses of 316L SS substrate, Cr, Co–29.3Cr and Co–29.1Cr/3TiO2
The presence of this peak, in spite of high thickness of this coating, coatings.
154 S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157

Fig. 7. Coefficient of friction for different samples during sliding. Fig. 9. SEM micrograph of the worn surface of the 316L SS substrate.

SS (Fig. 2). This discrepancy can be explained by considering this The friction coefficients of the substrate, Co–Cr and Co–Cr/TiO2
fact that the Archard’s theory is limited to idealized sliding condi- increase in the early stages of the sliding and then remain at the
tions. In this theory, the processes of crack nucleation and subse- same levels to the end of the tests. Average coefficients of friction
quent growth are disregarded [49]. Toughness also plays an for 316L SS, Cr, Co–Cr and Co–Cr/TiO2 are 0.86, 0.81, 0.70 and 0.61,
important role in cracking and wear resistance [28]. The wear respectively. The friction force in tribosystems usually results from
behavior of the chromium electrodeposits is a combined result of the adhesion between counterparts and ploughing work [5,50].
hard, brittle characteristics and cracked morphology of these coat- Higher friction coefficient of 316L SS, as compared to the alloy
ings [11]. Therefore, low wear resistance of this coating can be due and composite films, is consistent with the wear results. This is
to high residual stresses and brittle nature of the electrodeposited relevant to the low hardness of the substrate (Fig. 2), resulting in
Cr. The presence of a great number of microcracks in this film the increased real contact area between the alumina ball and
results in easier growth of cracks during wear and higher wear 316L SS substrate.
losses [9]. Internal stresses and microcracks densities are In the early stages of the wear test of the electrodeposited Cr,
decreased by co-deposition of Co with Cr and production of Co– the coefficient of friction rises at first and then decreases. This
Cr alloy coating. Better wear resistance of Co–Cr alloy and compos- trend may be a result of nodular surface morphology of Cr film.
ite films, despite their lower hardness as compared to the Cr coat- These nodules increase surface roughness and can act as an obsta-
ing, is because of their lower internal stresses along with improved cle against alumina ball at low sliding distances. The coefficient of
toughness and ductility. friction decreases after decapitation of the nodules. After this stage,
the coefficient of friction gradually increases up until the end of the
3.3.2. Coefficient of friction test, which could be due to the increased real contact area and
During the wear tests, the lateral forces were measured and detachment of sections of the Cr film. These peeled off sections also
converted to the coefficient of friction. The coefficients of friction act as the hard third body abrasive and hence increase the coeffi-
for three different coatings and the substrate are shown in Fig. 7. cient of friction [49].

Fig. 8. Wear track morphology of (a) 316L SS substrate, (b) Cr, (c) Co–Cr and (d) Co–Cr/TiO2 coatings.
S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157 155

Fig. 10. SEM micrographs from the (a and c) worn surface and (b and d) cross-section of Cr coating. (a) Secondary electron and (b–d) backscattered electron images.

3.3.3. Wear mechanism


Typical SEM micrographs of worn surfaces of the 316L SS sub-
strate, Cr, Co–Cr, and Co–Cr/TiO2 are shown in Fig. 8. Wear track
of the substrate is wider than the other samples. Narrowest wear
track is observed in the case of Co–Cr/TiO2 coating. These results
are in agreement with the volume loss values (Fig. 6). However,
in spite of relatively narrow wear path of Cr coating, detachment
of chromium in some areas increases volume losses, as shown in
Fig. 8(b).
Continuous grooves on the worn surface of the 316L SS sub-
strate are clearly observed in Fig. 8(a), indicating abrasion as the
prominent wear mechanism in this sample. These grooves are pro-
duced by micro-ploughing action of the hard asperities on the alu-
mina ball [3,5,50]. Higher magnification SEM micrograph reveals
that limited adhesion and plastic flow in a direction perpendicular
to the sliding direction also occurs during sliding (Fig. 9).
Fig. 10(a) shows deep grooves and severe disruption on the
worn surface of Cr coating. As the Cr coating is more brittle than
the 316L SS, the grooves on the worn surface of this film can be
due to the micro-cutting action. Micro-cutting is more dangerous
than micro-ploughing, which results in the higher volume losses
owing to the formation and detachment of chip from the surface
[3]. On the other hand, it is clear from this figure that a part of
the coating has been peeled off as a result of cracks propagation.
Cr electrodeposits are brittle and contain microcracks. During the
wear test, these microcracks propagate and grow under the sur-
face. A part of Cr coating is peeled off after two subsurface cracks
come together. Propagation of cracks under the surface of Cr coat-
ing is clear in the cross-section image (Fig. 10(b)). Fig. 10(b) reveals
that Cr electrodeposit has failed in a brittle manner. The detached
Cr debris also act as the hard third body abrasives which in turn
increase wear losses and the friction coefficients [11,28,49]. Debris Fig. 11. High magnification SEM micrographs from the worn surfaces of (a) Co–Cr
can oxidize during wear process and form islands on the wearing and (b) Co–Cr/TiO2 coatings.
surface. A Backscattered electron micrograph from the same region
of Fig. 10(a) is presented in Fig. 10(c). According to EDS analysis, coefficient of friction usually decrease after formation of oxide lay-
black regions in this figure are oxidized chromium and contain ers [3]. However, wear resistance of Cr electrodeposit is not
about 5.2, 42.3 and 52.5 at.% Al, Cr and O, respectively. The pres- improved by the formation of non-continuous oxide films on the
ence of Al in the EDS analysis can be due to the alumina ball wear- worn surface. In addition, as propagation of cracks along the inter-
ing by hard chromium debris. Formation of oxidized layer is also face between the oxide layer and the Cr coating is also clearly
observed in transverse section (Fig. 10(d)). Wear volume loss and observed in Fig. 10(d), the formed oxide films are easily detached
156 S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157

from the wearing surface and leave the fresh Cr surface in contact [2] H.F. Xuan, Q.Y. Wang, S.L. Bai, Z.D. Liu, H.G. Sun, P.C. Yan, A study on
microstructure and flame erosion mechanism of a graded Ni–Cr–B–Si coating
with the alumina ball.
prepared by laser cladding, Surf. Coat. Technol. 244 (2014) 203–209.
The wear behavior of Co–Cr and Co–Cr/TiO2 coatings is different [3] G. Bolelli, L. Lusvarghi, Heat treatment effects on the tribological performance
to that of the electrodeposited Cr, because they are softer and more of HVOF sprayed Co–Mo–Cr–Si coatings, J. Therm. Spray Technol. 15 (2006)
ductile than Cr film. High magnification SEM micrographs from the 802–810.
[4] C.U. Chisholm, Cobalt-chromium coatings by electrodeposition: review and
worn surfaces of the Co–Cr and nano-composite coatings are initial experimental studies, Electrodepos. Surf. Treat. 3 (1975) 321–333.
shown in Fig. 11. The worn surfaces of these coatings are relatively [5] S. Mahdavi, S.R. Allahkaram, Characteristics of electrodeposited cobalt and
smooth, exhibiting some shallow grooves. Flattened nodules are titania nano-reinforced cobalt composite coatings, Surf. Coat. Technol. 232
(2013) 198–203.
also observed on the worn surface of the composite film. Unlike [6] M. Srivastava, V.K. William Grips, K.S. Rajam, Electrochemical deposition and
316L SS substrate, no plastic deformation is observed on the worn tribological behaviour of Ni and Ni–Co metal matrix composites with SiC nano-
surfaces, because the load bearing capacity of materials is particles, Appl. Surf. Sci. 253 (2007) 3814–3824.
[7] G. Wu, N. Li, D. Zhou, K. Mitsuo, Electrodeposited Co–Ni–Al2O3 composite
increased by improving their hardness and mechanical properties coatings, Surf. Coat. Technol. 176 (2004) 157–164.
[3,50]. Therefore, mild abrasive wear is the main wear mechanism [8] M. Ebrahim-Ghajari, S.R. Allahkaram, S. Mahdavi, Corrosion behaviour of
in these coatings, which is in accordance to the obtained hardness electrodeposited nanocrystalline Co and Co/ZrO2 nanocomposite coatings,
Surf. Eng. 31 (2015) 251–257.
and volume losses (Figs. 2 and 6). [9] M. Rezaei-Sameti, S. Nadali, J. Rajabi, M. Rakhshi, The effects of pulse
electrodeposition parameters on morphology, hardness and wear behavior of
nano-structure Cr–WC composite coatings, J. Mol. Struct. 1020 (2012) 23–27.
4. Conclusions [10] S. Yagi, K. Murase, T. Hirato, Y. Awakura, Alternating pulsed electrolysis for
iron-chromium alloy coatings with continuous composition gradient, J.
In the present study, Cr, Co–Cr and Co–Cr/TiO2 nano-composite Electrochem. Soc. 154 (2007) 304–309.
[11] D.P. Weston, P.H. Shipway, S.J. Harris, M.K. Cheng, Friction and sliding wear
coatings were produced from Cr(III) based electrolytes, and their behaviour of electrodeposited cobalt and cobalt–tungsten alloy coatings for
composition, hardness, morphology and microstructure, together replacement of electrodeposited chromium, Wear 267 (2009) 934–943.
with tribological behavior were examined and compared to the [12] N.A. Polyakov, Y.M. Polukarov, V.N. Kudryavtsev, Electrodeposition of
composite chromium coatings from Cr(III) sulfate–oxalate solution
316L SS substrate. The outcome of the results can be summarized
suspensions containing Al2O3, SiC, Nb2N, and Ta2N particles, Prot. Met. Phys.
as follows: Chem. Surf. 46 (2010) 75–81.
[13] V.S. Protsenko, F.I. Danilov, Chromium electroplating from trivalent chromium
1. An increase in the concentration of TiO2 nano-particles concen- baths as an environmentally friendly alternative to hazardous hexavalent
chromium baths: comparative study on advantages and disadvantages, Clean
tration from 0 to 30 g L 1 in the bath could result in the Technol. Environ. Policy 16 (2014) 1201–1206.
increased amounts of their co-deposition and Cr content of [14] R. Giovanardi, G. Orlando, Chromium electrodeposition from Cr(III) aqueous
the coatings from 0 to 3.5 wt.% and from 23.7 to 31.2 wt.%, solutions, Surf. Coat. Technol. 205 (2011) 3947–3955.
[15] B. liu, Z. Zeng, Y. Lin, Mechanical properties of hard Cr–MWNT composite
respectively. coatings, Surf. Coat. Technol. 203 (2009) 3610–3613.
2. All the coatings indicated higher microhardness values as com- [16] C.A. Huang, C.K. Lin, C.Y. Chen, Hardness variation and corrosion behavior of
pared to that of the 316L SS substrate. Microhardness of Cr coat- as-plated and annealed Cr–Ni alloy deposits electroplated in a trivalent
chromium-based bath, Surf. Coat. Technol. 203 (2009) 3686–3691.
ing was about 920 Hv which was higher than that of Co–Cr and [17] Y.I. Choi, T. Eguchi, T. Asao, K. Kuroda, M. Okido, Mechanism for the formation
even Co–Cr/TiO2 films. Microhardness of Co–29.3Cr alloy coat- of black Cr–Co electrodeposits from Cr3+ solution containing oxalic acid, J.
ing was about 6.0% higher than Co–23.7Cr sample. Microhard- Electrochem. Soc. 161 (2014) 713–718.
 nienė, A. Lisowska-Oleksiak, The use of XPS
[18] S. Survilienė, V. Jasulaitienė, A. Češu
ness of the alloy coatings improved with co-deposition of up for study of the surface layers of Cr–Co alloy electrodeposited from Cr(III)
to 3 wt.% TiO2 nano-particles to a value of 805 Hv and from then formate–urea baths, Solid State Ionics 179 (2008) 222–227.
on remained almost constant. [19] T. Ohgai, Y. Tanaka, T. Fujimaru, Soft magnetic properties of Ni–Cr and Co–Cr
alloy thin films electrodeposited from aqueous solutions containing trivalent
3. All the coatings exhibited nodular morphology. Nodules size
chromium ions and glycine, J. Appl. Electrochem. 42 (2012) 893–899.
and micro-cracks densities of Co–Cr and Co–Cr/TiO2 coatings [20] C.A. Huang, U.W. Lieu, C.H. Chuang, Role of nickel undercoat and reduction-
were lower than those of Cr film. Electrodeposited Cr coating flame heating on the mechanical properties of Cr–C deposit electroplated from
had an amorphous structure while alloy and composite films a trivalent chromium based bath, Surf. Coat. Technol. 203 (2009) 2921–2926.
[21] Z. Zeng, J. Zhang, Electrodeposition and tribological behavior of amorphous
contained solid solutions of Cr in the hcp lattice of Co structure. chromium-alumina composite coatings, Surf. Coat. Technol. 202 (2008) 2725–
4. The wear resistance of the 316L SS substrate was improved by 2730.
about 3 and 7.5 times after electrodeposition of Co–29.3Cr and [22] B. Li, A. Lin, X. Wu, Y. Zhang, F. Gan, Electrodeposition and characterization of
Fe–Cr–P amorphous alloys from trivalent chromium sulfate electrolyte, J.
Co–29.1Cr/3.0TiO2 coatings, respectively. The wear volume loss Alloys Comp. 453 (2008) 93–101.
of Cr coating was about 13% higher than that of the substrate. [23] G. Saravanan, S. Mohan, Structure, composition and corrosion resistance
5. The composite coating had the lowest coefficient of friction. The studies of Co–Cr alloy electrodeposited from deep eutectic solvent (DES), J.
Alloys Comp. 522 (2012) 162–166.
friction coefficient of the substrate, Co–29.3Cr and Co–29.1Cr/ [24] B.A. Spiridonov, Electroplating of chromium-cobalt alloy coatings from sulfate
3.0TiO2 samples increased in the early stages of the sliding solutions, Prot. Met. 41 (2005) 249–253.
and then remained almost constant to the end of the test. A dif- [25] M.S. Marwah, V. Srinivas, A.K. Pandey, S.R. Kumar, K. Biswas, J. Maity,
Morphological changes during annealing of electrodeposited Ni–Cr coating on
ferent trend had been observed in Cr electrodeposit. steel and their effect on corrosion in 3% of NaCl solution, J. Iron Steel Res. Int.
6. Mild abrasive wear was the prominent wear mechanism in Co– 18 (2011) 72–78.
29.3Cr and Co–29.1Cr/3.0TiO2 coatings. Wear grooves were [26] E.A. Efimov, V.V. Chernykh, Electroplating with chromium–cobalt alloy, Prot.
Met. 37 (2001) 396–397.
deeper on the worn surface of the substrate. Limited plastic
[27] D. Cheng, V.L. Tellkamp, C.J. Lavernia, E.J. Lavernia, Corrosion properties of
deformation and adhesion were also observed on the worn sur- nanocrystalline Co–Cr coatings, Ann. Biomed. Eng. 29 (2001) 803–809.
face of this sample. Crack propagation and detachment of Cr [28] R. Ahmed, H.L. de Villiers Lovelock, S. Davies, N.H. Faisal, Influence of Re-
were considered to be the main reasons for high volume loss HIPing on the structure–property relationships of cobalt-based alloys, Tribol.
Int. 57 (2013) 8–21.
of this sample. [29] M.L. Lau, E. Strock, A. Fabel, C.J. Lavernia, E.J. Lavernia, Synthesis and
characterization of nanacrystalline Co–Cr coatings by plasma spraying,
Nanostruct. Mater. 10 (1998) 723–730.
[30] A. Ozols, M. Barreiro, E. Forlerer, H.R. Sirkin, Coating of Co–Cr–Mo alloy for
References surgical implants by centrifugal spray: preliminary evaluation, Surf. Coat.
Technol. 200 (2006) 5884–5888.
[31] B.V. Krishna, W. Xue, S. Bose, A. Bandyopadhyay, Functionally graded
[1] R. Li, T. Yuan, Z. Qiu, A model to describe the surface gradient-nanograin
Co–Cr–Mo coating on Ti–6Al–4V alloy structures, Acta Biomater. 4 (2008)
formation and property of friction stir processed laser Co–Cr–Ni–Mo alloy,
697–706.
Appl. Surf. Sci. 308 (2014) 176–183.
S. Mahdavi, S.R. Allahkaram / Journal of Alloys and Compounds 635 (2015) 150–157 157

[32] K. Kumagai, N. Nomura, T. Ono, M. Hotta, A. Chiba, Dry friction and wear [42] Y. Yu-fang, G. Zhu-qing, D. Li-yuan, L. Bei-ping, M. Yu-tian, Y. Zhen-hui,
behavior of forged Co–29Cr–6Mo alloy without Ni and C additions for implant Electrodeposition of Ni–Cr alloy on aluminum substrate, J. Cent. South Univ.
applications, Mater. Trans. 46 (2005) 1578–1587. Technol. 13 (2006) 219–224.
[33] I. Czako-Nagy, M.K. El-Sharif, A. Vertes, C.U. Chisholm, Studies of [43] M. Adabi, A.A. Amadeh, Electrodeposition mechanism of Ni–Al composite
electrodeposited chromium-cobalt alloy coatings by emission Co-57 coating, Trans. Nonferrous Met. Soc. China 24 (2014) 3189–3195.
Mossbauer spectroscopy, Electrochim. Acta 39 (1994) 801–805. [44] C.D. Gómez-Esparza, J. Camarillo-Cisneros, I. Estrada-Guel, J.G. Cabañas-
[34] P.C. Tulio, I.A. Carlos, Effect of SiC and Al2O3 particles on the electrodeposition Moreno, J.M. Herrera-Ramírez, R. Martínez-Sánchez, Effect of Cr, Mo and Ti
of Zn, Co and ZnCo: II. Electrodeposition in the presence of SiC and Al2O3 and on the microstructure and vickers hardness of multi-component systems, J.
production of ZnCo–SiC and ZnCo–Al2O3 coatings, J. Appl. Electrochem. 39 Alloys Comp. 615 (2014) S638–S644.
(2009) 1305–1311. [45] A. Neville, V. Kollia-Rafailidi, A comparison of boundary wear film formation
[35] B. Bahadormanesh, A. Dolati, M.R. Ahmadi, Electrodeposition and on steel and a thermal sprayed Co/Cr/Mo coating under sliding conditions,
characterization of Ni–Co/SiC nanocomposite coatings, J. Alloys Comp. 509 Wear 252 (2002) 227–239.
(2011) 9406–9412. [46] M. Masoudi, M. Hashim, H. Mohamed Kamari, M. Sapuan Salit, Fabrication and
[36] M. Li, Z. Ke-chao, L. Zhi-you, W. Qiu-ping, Electrodeposition of Ni–Co–Fe2O3 characterization of Ni–SiC–Cr nanocomposite coatings, Appl. Nanosci. 3 (2013)
composite coatings, J. Cent. South Univ. Technol. 17 (2010) 708–714. 357–362.
[37] S. Surviliene, V. Jasulaitiene, A. Lisowska-oleksiak, V.A. Safonov, Effect of WC [47] J.F. Archard, Contact and rubbing of flat surfaces, J. Appl. Phys. 24 (1953) 981–
on electrodeposition and corrosion behaviour of chromium coatings, J. Appl. 988.
Electrochem. 35 (2005) 9–15. [48] X.J. Sun, J.G. Li, Friction and wear properties of electrodeposited nickel–titania
[38] J. Gao, J. Suo, Preparation and characterization of the electrodeposited Cr– nanocomposite coatings, Tribol. Lett. 28 (2007) 223–228.
Al2O3/SiC composite coating, Appl. Surf. Sci. 257 (2011) 9643–9648. [49] S. Mahdavi, F. Akhlaghi, Effect of the SiC particle size on the dry sliding wear
[39] J. Valant, D. Drobne, Biological reactivity of TiO2 nanoparticles assessed by behavior of SiC and SiC–Gr-reinforced Al6061 composites, J. Mater. Sci. 46
ex vivo testing, Protoplasma 249 (2012) 835–842. (2011) 7883–7894.
[40] N. Guglielmi, Kinetics of the deposition of inert particles from electrolytic [50] S. Mahdavi, F. Akhlaghi, Effect of the graphite content on the tribological
baths, J. Electrochem. Soc. 119 (1972) 1009–1012. behavior of Al/Gr and Al/30SiC/Gr composites processed by in situ powder
[41] B.A. Spiridonov, Electrodeposition of chromium-cobalt alloy coatings from metallurgy (IPM) method, Tribol. Lett. 44 (2011) 1–12.
sulfate electrolytes, Russ. J. Appl. Chem. 76 (2003) 65–70.

You might also like