11
The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings C. Ma a , S.C. Wang a, , L.P. Wang b , F.C. Walsh a,c , R.J.K. Wood a a National Centre for Advanced Tribology at Southampton (nCATS), Engineering Sciences, University of Southampton, Higheld, Southampton SO17 1BJ, UK b Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou, PR China c Engineering Materials, Engineering Sciences, University of Southampton, Higheld, Southampton SO17 1BJ, UK abstract article info Article history: Received 17 March 2013 Accepted in revised form 4 August 2013 Available online xxxx Keywords: Co-Ni-P Friction Microstructure Heat treatment Wear A new Co-Ni-P coating having superior properties to hard chrome coatings for anti-wear applications has been developed by combining the precipitation hardening of Ni-P alloys with the lubricity of cobalt-rich Co-Ni coatings. The evolution of composition and microstructure, hardness, thermal stability and tribological properties have been investigated. The local pH value near the cathode played an important role in the change of the microstructure from nanocrystalline to amorphous along the growth direction as the phosphorous content increased from 7 at.% to 26 at.%. The highest microhardness (980 HV) and the lowest wear rate (one order of magnitude lower than that of hard chrome coatings under the same dry sliding conditions) were achieved by annealing the coatings at 400 °C facilitating precipitation hardening. Furthermore, the coefcients of friction of both the as-deposited Co-Ni-P coating and the heat-treated samples were approximately 0.3, only half of that of hard chrome coatings. The wear rates of Co-Ni-P coatings annealed up to 400 °C are 47 times lower compared to that of hard chrome. The roll-like debris found on the worn surfaces of the coating annealed at 500 °C were oriented perpendicularly to the sliding direction. Such part might be related to the slightly reduced coefcient of friction in the stable stage. © 2013 Elsevier B.V. All rights reserved. 1. Introduction For over a century, hard chrome coatings have been applied in a wide range of industries but are facing strict regulations due to both work safety and environmental issues. The chrome plating solution consists of chromic acid with hexavalent Cr, which is extremely toxic and carcinogenic [1,2]. Substitutable materials, alternative technologies and new designs have attracted increasing scientic and industrial attention, including non-Cr electroplating, electroless plating, trivalent Cr plating, high velocity oxy-fuel thermal spraying and physical vapour deposition [3,4]. Among these current methods, electrodeposition is versatile, technologically feasible and economically competitive. Ni-P and its composite coatings, such as Ni-P-Al 2 O 3 , Ni-P-SiC and Ni-P-B 4 C, have been considered as alternatives to hard chrome coatings for anti- wear applications [2,5]. It has been reported that the electroplated Ni-P coatings have high hardness (more than 640 HV) and improved wear resistance [6]. The hardness could further reach 1000 HV after heat treatment at 400 °C due to the precipitation strengthening, which resulted in a lower wear rate than that of hard chrome coatings [2]. However, the coefcient of friction of Ni-P is in the range of 0.450.7 on dry sliding conditions against different counterface materials, including carbon steel, TiN-coated tool steel and Si 3 N 4 [2,7,8]. Wang et al. [9] and Ma et al. [10] reported that electrodeposited cobalt-rich Ni-Co alloys exhibited a low coefcient of friction (lower than 0.3) and enhanced wear resistance. By alloying Co into Ni-P, Co-Ni-P coatings are expected to exhibit a lower coefcient of fric- tion and improved wear resistance. Thin Co-Ni-P lms (from 10 nm to 10 μm thick) are generally used as magnetic recording media, and as hard magnetic materials for microelectromechanical systems applications due to their unique mag- netic properties [11,12]. In the engineering industry, they have been ap- plied as anti-corrosion coatings [13], microwave absorptive materials [14] and electrocatalytic materials for water electrolysis [15]. However, few researchers have addressed the tribological properties of the Co-Ni-P coatings. In the present research, we aim to combine the precipitation hardening found in Ni-P alloys and the lubricity of cobalt-rich Co-Ni coatings to fabricate a novel friction reducing and wear-resistant Co-Ni-P coating to replace hard chrome. Co-Ni-P lms have been developed by electroless plating [12,16,17] and electodeposition [13,1821]. Compared to electroless plating, the versatile low-cost electrodeposition has a faster plating rate and more stable baths. Co-Ni-P lms can be electroplated from a sulphate bath [13], a Brenner type chloride bath [19,20] or a sulphamate bath [21] on copper or mild steel substrates. The experimental parameters, such as pH, current density and electrolyte composi- tions, have been reported to affect the surface morphology, composi- tion, microstructure, microhardness and their magnetic properties. Pattanaik et al. [19] reported that the Co-Ni-P lm with the P content more than 12 at.% had an amorphous structure. Djokić et al. [22] Surface & Coatings Technology xxx (2013) xxxxxx Corresponding author. Tel.: +44 23 8059 4638; fax: +44 23 8059 3016. E-mail address: [email protected] (S.C. Wang). SCT-18792; No of Pages 11 0257-8972/$ see front matter © 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.surfcoat.2013.08.009 Contents lists available at ScienceDirect Surface & Coatings Technology journal homepage: www.elsevier.com/locate/surfcoat Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://dx.doi.org/10.1016/j.surfcoat.2013.08.009

The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

  • Upload
    rjk

  • View
    215

  • Download
    2

Embed Size (px)

Citation preview

Page 1: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Surface & Coatings Technology xxx (2013) xxx–xxx

SCT-18792; No of Pages 11

Contents lists available at ScienceDirect

Surface & Coatings Technology

j ourna l homepage: www.e lsev ie r .com/ locate /sur fcoat

The electrodeposition and characterisation of low-friction andwear-resistant Co-Ni-P coatings

C. Ma a, S.C. Wang a,⁎, L.P. Wang b, F.C. Walsh a,c, R.J.K. Wood a

a National Centre for Advanced Tribology at Southampton (nCATS), Engineering Sciences, University of Southampton, Highfield, Southampton SO17 1BJ, UKb Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou, PR Chinac Engineering Materials, Engineering Sciences, University of Southampton, Highfield, Southampton SO17 1BJ, UK

⁎ Corresponding author. Tel.: +44 23 8059 4638; fax: +E-mail address: [email protected] (S.C. Wang).

0257-8972/$ – see front matter © 2013 Elsevier B.V. All rihttp://dx.doi.org/10.1016/j.surfcoat.2013.08.009

Please cite this article as: C. Ma, et al., Surf. C

a b s t r a c t

a r t i c l e i n f o

Article history:Received 17 March 2013Accepted in revised form 4 August 2013Available online xxxx

Keywords:Co-Ni-PFrictionMicrostructureHeat treatmentWear

A new Co-Ni-P coating having superior properties to hard chrome coatings for anti-wear applications has beendeveloped by combining the precipitation hardening of Ni-P alloys with the lubricity of cobalt-rich Co-Ni coatings.The evolution of composition andmicrostructure, hardness, thermal stability and tribological properties have beeninvestigated. The local pH value near the cathode played an important role in the change of the microstructurefrom nanocrystalline to amorphous along the growth direction as the phosphorous content increased from7 at.% to 26 at.%. The highest microhardness (980 HV) and the lowest wear rate (one order of magnitude lowerthan that of hard chrome coatings under the samedry sliding conditions)were achieved by annealing the coatingsat 400 °C facilitating precipitation hardening. Furthermore, the coefficients of friction of both the as-depositedCo-Ni-P coating and the heat-treated samples were approximately 0.3, only half of that of hard chrome coatings.The wear rates of Co-Ni-P coatings annealed up to 400 °C are 4–7 times lower compared to that of hard chrome.The roll-like debris found on the worn surfaces of the coating annealed at 500 °C were oriented perpendicularlyto the sliding direction. Suchpartmight be related to the slightly reduced coefficient of friction in the stable stage.

© 2013 Elsevier B.V. All rights reserved.

1. Introduction

For over a century, hard chrome coatings have been applied in awide range of industries but are facing strict regulations due to bothwork safety and environmental issues. The chrome plating solutionconsists of chromic acid with hexavalent Cr, which is extremely toxicand carcinogenic [1,2]. Substitutable materials, alternative technologiesand new designs have attracted increasing scientific and industrialattention, including non-Cr electroplating, electroless plating, trivalentCr plating, high velocity oxy-fuel thermal spraying and physical vapourdeposition [3,4]. Among these current methods, electrodeposition isversatile, technologically feasible and economically competitive. Ni-Pand its composite coatings, such as Ni-P-Al2O3, Ni-P-SiC and Ni-P-B4C,have been considered as alternatives to hard chrome coatings for anti-wear applications [2,5]. It has been reported that the electroplatedNi-P coatings have high hardness (more than 640 HV) and improvedwear resistance [6]. The hardness could further reach 1000 HV afterheat treatment at 400 °C due to the precipitation strengthening,which resulted in a lower wear rate than that of hard chrome coatings[2]. However, the coefficient of friction of Ni-P is in the range of0.45–0.7 on dry sliding conditions against different counterfacematerials, including carbon steel, TiN-coated tool steel and Si3N4

[2,7,8]. Wang et al. [9] and Ma et al. [10] reported that electrodeposited

44 23 8059 3016.

ghts reserved.

oat. Technol. (2013), http://d

cobalt-rich Ni-Co alloys exhibited a low coefficient of friction (lowerthan 0.3) and enhanced wear resistance. By alloying Co into Ni-P,Co-Ni-P coatings are expected to exhibit a lower coefficient of fric-tion and improved wear resistance.

Thin Co-Ni-P films (from 10 nm to 10 μm thick) are generally usedas magnetic recording media, and as hard magnetic materials formicroelectromechanical systems applications due to their unique mag-netic properties [11,12]. In the engineering industry, they have been ap-plied as anti-corrosion coatings [13], microwave absorptive materials[14] and electrocatalytic materials for water electrolysis [15]. However,few researchers have addressed the tribological properties of the Co-Ni-Pcoatings. In the present research, we aim to combine the precipitationhardening found in Ni-P alloys and the lubricity of cobalt-rich Co-Nicoatings to fabricate a novel friction reducing and wear-resistantCo-Ni-P coating to replace hard chrome.

Co-Ni-P films have been developed by electroless plating [12,16,17]and electodeposition [13,18–21]. Compared to electroless plating, theversatile low-cost electrodeposition has a faster plating rate and morestable baths. Co-Ni-P films can be electroplated from a sulphatebath [13], a Brenner type chloride bath [19,20] or a sulphamatebath [21] on copper or mild steel substrates. The experimentalparameters, such as pH, current density and electrolyte composi-tions, have been reported to affect the surface morphology, composi-tion, microstructure, microhardness and their magnetic properties.Pattanaik et al. [19] reported that the Co-Ni-P film with the P contentmore than 12 at.% had an amorphous structure. Djokić et al. [22]

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 2: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 1. Evolution of (a) bulk solution pH (b) local pHvalue near the cathode as a function ofplating time.

2 C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

found that Co-Ni-P coatings exhibited an amorphous structure onthe top surface but a crystalline structure at the beginning of electro-deposition. It was suggested that the P content controlled the transi-tion. However, its mechanism is not clear.

In the present research, Co-Ni-P coatings have been electrodepositedfrom a modified sulphate bath. The effect of local pH value near thecathode on the microstructure and composition with increased platingtimeswere investigated. Furthermore, the hardness and the tribologicalproperties of Co-Ni-P coatings after heat treatment were evaluated andcompared with those of hard chrome coatings.

2. Experimental details

2.1. Electrodeposition

Nanocrystalline Co-Ni-P coatings were electrodeposited from theelectrolytes consisting of 40 g dm−3 NiSO4 · 6H2O, 40 g dm−3

CoSO4 · 7H2O, 20 g dm−3 NaH2PO2 and additives. The electrodeposi-tion was carried out with the current density of 5 A dm−2 at 45 °C.High purity Ni sheets (40 mm × 10 mm × 1 mm) were used as theanode and the cathode was AISI 1020 mild steel (40 mm × 10 mm ×3 mm). The solution was continuously stirred by a PTFE-coated mag-netic stirrer bar (6 mm diameter × 30 mm length) at 200 rpm. Theplating time was controlled from 1 min to 75 min. The pH values ofelectrolytic bath were measured before and after the electrodeposition.The in situ measurement of the local pH value near the cathode wascarried out by fixing the pH electrode as close to the cathode surfaceas possible. The as-deposited samples were plated for 75 min. In orderto study the influence of the substrate, the plating was repeated ontop of the previous Co-Ni-P coating. Electrochemical measurementswere carried out by an Autolab PGSTAT30 potentiostat/galvanostat sys-temwith a nickel counter electrode (40 mm × 10 mm × 1 mm) and asaturated calomel electrode (SCE) reference electrode. The as-depositedCo-Ni-P coatings were subsequently annealed in air at 200 °C, 300 °C,400 °C and 500 °C for 1 h, respectively. As a comparison, the hardchrome coatings were electrodeposited on similar mild steel substratesfrom a conventional chromic acid bath (CrO3 160 g dm−3, H2SO4

2.5 g dm−3) containing 2 g dm−3 CrCl3 with a direct current of 100 Adm−2 at 45 °C. In order to compare tribological properties, Ni-P coatingswith the same P content as the present Co-Ni-P coatings were platedfrom the sulphate bath as well.

2.2. Deposit morphology, composition and hardness

The surface morphology and the composition of the as-depositedand annealed Co-Ni-P coatings as well as their cross sections wereinvestigated by a JOEL JSM 6500 SEM equipped with EDS utilizing thesecondary and backscattered electron imaging (SEI and BEI). The sur-face roughness of the coating was measured by an Agilent 5500 atomicforce microscope (AFM). A line profile (5 μm) with height measure-ments was used to determine the surface roughness. Five measure-ments were performed on each coating. The crystal structure wascharacterised by XRD using Cu Kα radiation. The morphology changealong the growth direction was further studied by focused ion beam(FIB) technology (FIB 1540 XB CrossBeam, Leo/Zeiss). The hardness ofdeposits was measured by a Vicker's microhardness indenter under anapplied loadof 100 g for 15 s. An average offivemeasurementswas car-ried out on each coating.

2.3. Tribological evaluation

The tribological behaviour was tested on a reciprocating TE-77tribometer (Phoenix, UK) under the dry sliding conditions with relativehumidity of 40–50% at room temperature (25 °C) in air. An AISI-52100stainless steel ball (diameter 6 mm) with a hardness of 700 HV wasused as the counter body. The surface roughness (Ra) of the pin is

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

0.48 μm. The calibrated load was 14 N corresponding to initial Hertziancontact pressure of 1.6 GPa with the sliding frequency of 1 Hz. Thesliding stroke was 2.69 mm, while the average sliding speed was5.38 mm/s. The total sliding time is 25 min. The friction force wasmea-sured by a piezo electric transducer and recorded automatically duringthe tests. The coefficient of friction was calculated by dividing thefriction force by the normal load. Three tests were performed oneach sample, and the average dynamic friction for the whole testswas obtained. The line profile of wear tracks was measured by anAlicona InfiniteFocus Real3D surface profilometer after wear testing,which illustrated the shape and depth of wear tracks for calculatingthe wear volume. The average cross-sectional area was determinedby the profiles at five locations perpendicular to the wear track. Thewear volume was calculated by multiplying the worn cross-sectionalarea by the sliding stroke (i.e., the wear track length). Subsequently,the average wear rates were determined in order to evaluate the anti-wear properties. Additionally, morphology and composition of theworn surface and debris were studied by SEM and EDS.

3. Results and discussion

3.1. Electrodeposition of Co-Ni-P coatings

Fig. 1(a) shows that the pH value of the bulk solution decreased fromapproximately 2.9 to 1.9 during the electroplating process. The pH fellmuch faster during the first 20 min. The local pH measurements wereperformed in order to evaluate the pH shift at the electrode surface. Itis found in Fig. 1(b) that the pH immediately jumped from 2.9 to 10 atthe very beginning of the electrodeposition. In the following 12 min, itrapidly dropped to 7 and was found to be proportional to t1/2. Subse-quently, it gradually changed from 7 to 6 at an approximate linear rate.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 3: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

3C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

The electrodeposition of Co andNi from the sulphate solution aswellas the codeposition of phosphorous follows the reactions below [22]:

Ni2þ þ 2e− →Ni ð1Þ

Co2þ þ 2e− → Co ð2Þ

H2PO2− þ e− → Pþ 2OH− ð3Þ

Simultaneous hydrogen evolution occurs:

2Hþ þ 2e− →H2 ð4Þ

The cathodic reactions (3) and (4) lead to an increase of pH, whichexplains the sudden pH rise near the cathode shown in Fig. 1(b). Thefollowing pH decrease may be attributed to hypophosphite oxidationand oxygen evolution near the anode [23,24]:

H2PO2− þH2O→H2PO

−3 þ 2Hþ þ 2e− ð5Þ

2H2O→O2 þ 4Hþ þ 2e− ð6Þ

The positively charged hydrogen ions were generated at the anodeand simultaneously diffused through the electrolyte to the cathode. Itwas suggested that the rate of the proton production at the anode washigher than the rate of the hydroxyl radical production at the cathode.After 12 min, the relative generation rate between hydrogen ions andhydroxyl radicals changed and the local pH value was more stable. Thenet generation of hydrogen ions led to the decrease in the pH value ofthe bulk solution.

Fig. 2(a) shows the electrodeposition potential (EED)-time (t) de-pendence for the electrodeposition of the Co-Ni-P coating. The EED-t

Fig. 2. (a) Electrode potential vs. time curves during Co-Ni-P deposition with (b) the en

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

curve can be divided into three characteristic regions. In the initialregion, the potential rapidly decreased during the first 20 s as shownin Fig. 2(b), which corresponded to the process of creating initial nucle-us on the mild steel substrate. Similar phenomena have been reportedfor nickel [25] and copper [26] electrodeposition, followed by a plateaufor the remaining deposition time. However, in the present research,the potential did not converge to a steady-state value. Instead, in theregion II from 20 s to 12 min, the potential gradually increased dueto the decreasing pH value near the cathode. For the electrode reactionH2PO2

− þ e−→Pþ 2OH− , the potential is linearly related to ln H2PO2−½ �∗

OH−½ �2∗,

where H2PO2−½ �∗ and [OH−]∗ are the concentration of H2PO2

− andOH− next to the cathode [27]. If the concentration of H2PO2

− is assumedto be constant, it can be deduced that the mixed electrodepositionpotential EED will be proportional to pH, which has a linear dependenceon t1/2. EED was plotted against t1/2 as shown in Fig. 2(c) and the rela-tionship of EED ∝ t1/2 was found. After 12 min, the potential linearly in-creased from −1.31 V to −1.27 V. The potential-time transientchanged from EED ∝ t1/2 in stage II to EED ∝ t in the stage III due to thevaried pH, which indicated different structure and composition of thedeposits.

3.2. Microstructure and composition of as-deposited Co-Ni-P coatings

Fig. 3 shows the surface morphologies of as-deposited Co-Ni-P coat-ings at different deposition times varying from 1 min to 75 min. After1 min deposition (Fig. 3(a)), the surface of the substrate was coveredby spherulites consisting of hundreds of grains as shown in the insertedhigh resolution image. The grain sizewas very small but clearly resolvedat the average of 5 nm. The layer after 10 min plating (Fig. 3(b))consisted of nodular clusters. The grain sizemeasured from thehigh res-olution image was also approximately 5 nm. The gaps were found

larged image of the first 100 s and (c) the EED versus t1/2 plot from 20 s to 12 min.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 4: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 3. Surface morphologies of as-deposited Ni-Co-P coatings after plating times of (a) 1 min, (b) 10 min, (c) 20 min, (d) 30 min, (e) 60 min and (f) 75 min.

4 C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

among aggregations of the crystals. After 20 min, the coating surfacewas characterised by the adjacent smooth areaswith no obvious spher-ulites as shown in Fig. 3(c). There were a lot of small pits with the diam-eter less than 300 nm distributed along the boundary of the smoothareas. Most of the small pits were covered by the smooth layer afterplating for 30 min as shown in Fig. 3(d). The smooth area was muchbigger with unclear boundaries but with a few larger pits. With the fur-ther increase of plating time, no boundaries were observed as shown inFig. 3(e, f). The surface was smoother with shallower pits. The surfaceroughness (Ra) was approximately 12 nm.

The SEM cross-sectional view of the as-deposited Co-Ni-P sampleplated for 75 min is shown in Fig. 4(a). The thickness of the coating is25 ± 3 μm. The initial deposit with the thickness of 6 μm grew incolumns, which was confirmed by the secondary ion image of FIB inFig. 4(b). As shown in Fig. 4(b), when the thickness was more than6 μm, it suddenly changed to very fine structure, which correspondedto the smooth layer shown in Fig. 3(c). Thus, it can be concluded thatthe structure change occurred during the plating time between

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

10 min and 20 min. The growth of the columnar structure Fig. 4(b) iscorresponding to the region II as shown in Fig. 2(a, c).

The distribution of the elements along the thickness direction of theas-deposited Co-Ni-P coating is shown in Fig. 4(c). The cobalt contentgradually decreases from 82 at.%, close to the substrate-coating inter-face to 61 at.% on the outer layer of the coating. The nickel content re-mains around 10–13 at.%. The phosphorous content increases from7 at.% to 26 at.% with the increase of the coating thickness. The percent-age of the less noble cobalt deposited preferentially, which has been re-ported in electroplated Ni-Co coatings [9,10]. The anomalous depositioncan be explained by the formation and adsorption of Co(II) and Ni(II)hydroxide ions on the cathode surface [28]. The noble Ni deposition issuppressed in the presence of Co(OH)+ preferentially formed andadsorbed on the cathode. As the plating time increased, the trend ofthe anomalous deposition was weakened due to the pH value decreaseas shown in Fig. 1. The lower pH value inhibits the formation of themetal hydroxide ions, which reduces the difference between Ni(OH)+

and Co(OH)+. Hereby, the preferred Co deposition is weakened. The

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 5: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 4. (a) SEM cross-sectional view of the as-deposited Co-Ni-P coating plated for 75 minwith (b) the FIB image of the enlarged area, (c) distribution of Co, Ni and P content alongthe thickness direction.

Fig. 5. XRD patterns of as-deposited Co-Ni-P coatings after plating times of (a) 1 min,(b) 10 min, (c) 20 min, (d) 30 min, (e) 60 min and (f) 75 min.

5C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

concentration of Co2+ is reduced by approximately 5% after electro-plating for 75 min, which also slightly contributes to the decreasingratio of Co/Ni. The deposition of P is promoted because of the decreasingpH value according to Eq. (3).

XRD patterns of the as-deposited Co-Ni-P coatings with differentplating times from1 min to 75 min are shown in Fig. 5. The coatings de-posited for 1 min has a peak at 2 θ ≈ 44.4 ∘ (Fig. 5(a)), which is consis-tent with the spacing of (111) in face-centered cubic (fcc) structured Co(and/or Ni) and/or (002) hexagonal close-packed (hcp) structured Colattice. As the thickness increases after plating for 10 min, the intensityof the peak at 2 θ ≈ 44.4° is much higher (Fig. 5(b)). The grain size wascalculated according to the Scherrer equation [29]:

d ¼ 0:9λBcosθ

ð7Þ

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

where λ is thewavelength of the X-rays (0.154 nm for Cu Kα radiation)and θ is the Bragg diffraction angle. B is the full width at half maximumheight (FWHM) of the peak. The calculated grain size of the coatingsplated for 1 min and 10 min is 22 nm. The total XRD peak broadeningis attributed to bothmicrostrain and the grain size [30]. Due to the inev-itable lattice distortion in the present coatings, the actual grain sizeshould shift away from the calculated value (22 nm). Our recent studyon the Ni-Co coating showed that the grain size measured by XRD wasconsistent with the one obtained by TEM with the grain size just over10 nm [31]. However, it seemed that the XRD failed to determine thegrain size of a few nanometres for the coatings plated after 1 min and10 min as shown in the present research.

The layer plated for 20 min with 17 at.% P exhibits the broad peakcentered at approximately 44.5° as shown in Fig. 5(c), which corre-sponds to the amorphous structure. With the further increase of theplating time, the microstructure keeps amorphous (Fig. 5(d–f)). Thetransition from the nanocrystalline structure to the amorphous struc-ture occurred after plating for 12 min, which is corresponding to thetransition from region II to region III in Fig. 2(a). The P content increasedfrom initial 7 at.% after 1 min plating to 15 at.% at the transition stage(12 min plating) then to 26 at.% at the end of plating (75 min).Meanwhile, the pH values near the cathode decreased from 10 to 8then 6, respectively. The structure evolution could be due to the atomicmismatch. The Co-Ni solid solution has the atomic diameter of0.250 nm, whereas the solute P has much smaller one at 0.217 nm.The mismatch between two different atoms is 13.2%, which is equiva-lent to a distortion of 1.98% for the 15 at.% P embedded in the Co-Nisolid solution. It appears the Co-Ni matrix can accommodate around2% strain to hold the crystal structure. Above this criterion, Co-Ni-P al-loys will favour the amorphous structure to accommodate more soluteP. A FIB/TEM combination and simulation could be helpful for the clari-fication, whichwill be carried out in futurework. As shown in Fig. 6, it isfound that the Co-Ni-P film plated on the previous Co-Ni-P coatingswith the amorphous surface layer has the same morphology evolution.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 6: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 6. SEM cross-sectional view of the as-deposited Co-Ni-P coating on the previousCo-Ni-P coating.

6 C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

Therefore, the change of the microstructure cannot be attributed to theinfluence of the substrate. Manna et al. [23] reported that the pH had noeffect on themicrostructure of the electroless Ni-P-Fe coatings based on

Fig. 7. Surface morphologies of heat treated Co-Ni-P coatings a

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

the results of the coatings treated for 30 min, 60 min and 90 min. In thepresent research, the pH is found to play an important role in the surfacemorphology, composition and microstructure. The contradictory resultis, perhaps, due to their lack of deposition data during the first 30 min.

3.3. Thermal stability

Fig. 7 shows the surface morphologies of the Co-Ni-P coatings afterheat treatment at different temperatures for 1 h. Compared to the as-deposited sample shown in Fig. 3(f), there is no significant change inthe morphology of the coating annealed at 200 °C (Fig. 7(a)), whichhas similar composition as the as-deposited sample except a smallamount of oxygen as shown in Fig. 8. The oxidation of the coating sur-face starts from 300 °C, which corresponds to the formation of fibrousoxides shown in Fig. 7(b). With the increase of annealing temperatures,the content of oxygen increases,while the content of Ni and P in the sur-face drops to 0 at 500 °C. The top layer of the coating annealed at 500 °Cis totally covered by the fibrous cobalt oxides shown in Fig. 7(d).The surface roughness increases to approximately 60 nm. As shown inthe Fig. 9(a) and (b), the 1 μm-thick cobalt oxide layer is followed bythe layer depleted in Co, which indicates that the alloys undergo theexternal oxidation [32]. It had been reported that the oxide film ofamorphous Ni77P23 oxidised at 300 °C in air mainly consists of NiO be-cause the oxide was formed by the migration of Ni2+ ions along grainboundaries in the oxide film [33]. In the present Co-Ni-P ternary alloy,as the activation energy for the motion of Co2+ is lower than that ofthe Ni2+ [34], the major oxidation product of the top layer is cobaltoxide.

Fig. 10 shows the XRD patterns of the as-deposited Co-Ni-P coatingand the samples annealed at different temperatures. No phase transi-tion occurs for the sample annealed at 200 °C. A number of small

t (a) 200 °C, (b) 300 °C, (c) 400 °C and (d) 500 °C for 1 h.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 7: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 8. Evolution of the elemental content from the deposit surface with annealingtemperature.

Fig. 10.XRDpatterns of (a) as-depositedCo-Ni-P coatingsplated for 75 min and the coatingsannealed at (b) 200 °C, (c) 300 °C, (d) 400 °C, (e) 500 °C for 1 h.

7C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

peaks corresponding to Ni12P5 emerge as the annealing temperatureincreases to 300 °C. After annealing at 400 °C, new peaks appear inFig. 10(d), which is consistent with the spacing of (200) and (220) infcc structured CoO. As confirmed in Fig. 10(e), the intensity of the twopeaks of the sample annealed at 500 °C increases as the top layermainlyconsists of CoO.

Fig. 9. (a) BEI cross-sectional viewof the Co-Ni-P coating annealed at 500 °C, (b) distributionof Co, Ni and P content along the thickness direction.

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

3.4. Microhardness

The variation of hardness with annealing temperatures is shown inFig. 11. The as-deposited Co-Ni-P coating plated has the microhardnessof 602 ± 15 HV, which is lower than the as-deposited hard chromecoatings. As the heat treatment temperature reaches 400 °C, themicrohardness increases to 980 ± 65 HV, which is comparable to thatof hard chrome coatings, and then decreases to 650 ± 30 HV afterannealing at 500 °C. Although no literature addresses on the hardeningmechanism of Co-Ni-P coatings, the significant hardening has beenreported for Ni-P alloys annealed at 400 °C [5,35,36]. It has been acceptedthat the precipitation of dispersed hard intermetallic compounds (Ni12P5and Ni3P) in the nickel matrix which can impede the dislocation move-ment result in the increased hardness. Similarly, the formation of Ni12P5precipitates is responsible for the significant increase of hardness of theCo-Ni-P coating. The decrease of the hardness after annealing at 500 °Cis probably due to the formation of oxide layer, the precipitate coarseningand Ni12P5 intermediate phase transition.

3.5. Friction and wear behaviour

Fig. 12 presents the comparison of coefficient of friction between ahard chromium coating and Co-Ni-P coating annealed at different tem-peratures under dry sliding wear conditions. The coefficients of frictionof both the as-deposited Co-Ni-P coating and the heat-treated samplesare approximately 0.3 and approximately half that of hard chrome coat-ings.Without cobalt, the Ni-P coatingwith the same P content (25 at.%)has higher coefficient of friction (0.5) under the same testing conditions.It is found that the presence of cobalt reduces the coefficient of friction.Typical curves of coefficient of friction of the as-deposited Co-Ni-P

Fig. 11. Variation of hardness of the coatings annealed at different temperatures for 1 hcompared with hard chrome coatings.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 8: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 12. Comparison of coefficient of friction between Co-Ni-P coatings annealed atdifferent temperatures and as-deposited hard chrome coatings under dry slidingwear conditions.

Fig. 14. Comparison ofwear rates between Co-Ni-P coatings annealed at different temper-atures and as-deposited hard chrome coatings under dry sliding wear conditions.

8 C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

coating and the coating annealed at 500 °C are shown in Fig. 13. The dif-ference between the two curves is in the period of the run-in stage,where the curve of the coating annealed at 500 °C is less stable thanits counterpart. It is attributed to the formation of cobalt oxide layerwith higher surface roughness as shown in Fig. 7(d). In the steady-stable stage, the coefficient of friction of the sample annealed at500 °C is slightly lower than that of the as-deposited sample.

Fig. 14 shows the effect of heat treatment temperature on the wearrates of Co-Ni-P coatings compared with hard chrome coatings. Thereis no significant change in the wear rate of the as-deposited coatingand the sample annealed at 200 °C because of their similar surfacemor-phology, composition, microstructure and hardness. As the annealingtemperature rises to 300 °C, the wear rate starts to decrease and thelowest wear rate (2.8 × 10−6 mm3N−1 m−1) is obtained after theheat treatment at 400 °C due to the precipitation hardening by Ni12P5precipitates. The wear rates of Co-Ni-P coatings annealed up to 400 °Care 4–7 times lower than that of hard chrome coatings. However, withfurther increase of the temperature to 500 °C the wear rate increasesto 1.3 × 10−5 mm3N−1 m−1.

The worn surfaces of the samples are shown in Fig. 15(a, c, e), indi-cating mild wear. The grooves and a small amount of wear debrisalong the sliding direction can be found on the worn track of the as-deposited sample (Fig. 15(a)). Theworn surface of the sample annealedat 400 °C shown in Fig. 15(c) is much smoother with less plastic

Fig. 13. The coefficient of friction of the as-deposited Co-Ni-P coating and the coatingannealed at 500 °C for 1 h.

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

deformation due to its high hardness. The cracks normal to the slidingdirection are observed on its worn surface. The wear debris is muchless than that of the as-deposited sample. When the annealing temper-ature increases to 500 °C, wider cracks appear on the worn surface(Fig. 15(e)), which results in its higher wear rate. The debris is morethan that of the sample annealed at 400 °C. The formation of cracks onthe worn surface can be attributed to the increase of the brittlenessand internal stress. It has been reported that the higher internal stressresulted in lower ductility and adhesion between coatings and sub-strates [37]. The effect of internal stress will be investigated in futurework. The pin wear scars were also examined by SEM and EDS. Thesize of the pin wear scar was in accordance with the wear rate. Thewear debris were attached to the pin surface.

Typical morphologies of the debris are shown in Fig. 15(b, d, f) whiletheir composition is listed in Table 1. There are two types of oxidised de-bris randomly distributed on thewear track of the as-deposited coatings(Fig. 15(b)), irregular shaped debris and the roll-like debris highlightedin the circles. The rolls are approximately 150–300 nm in diameter, andthe length varies from 0.35 μm to 2.6 μm. As pointed by the arrows, theplate-like flakes start to be detached from the surface and the edge ofthe flakes has been warped due to the residual stress [38]. As the heattreatment temperature increases to 400 °C, the plate-like flakes whichare about to peel off are more obvious as the worn surface is smoother(Fig. 15(d)). Unlike the as-deposited coating, the debris of the sampleannealed at 400 °C contains 2 at.% Fe from the pin. It can be attributedto its high hardness. It is interesting to find that most of the roll-like de-bris of the coating annealed at 500 °C shown in Fig. 15(f) are perpendic-ular to the sliding direction. It seemed that the flakes detached from thesurface of the coating annealed at 500 °C were easier to roll up thanthose of the other two coatings, whichmay be attributed to its differentsurface composition. The total cobalt oxide layer followed by the layerdepleted in Co may introduce more residual stress to form the rolls,the morphology of which is shown in the enlarged view in Fig. 15(f).The closer inspection reveals that the inner part of the roll-like debrisis darker than the edge, which indicates that the rolls are not solid. Dur-ing the sliding test, the easily formed rolls tend to be aligned normal tothe sliding direction. No iron is detected in the debris.

The roll-like debris have been reported by several researchers in ce-ramics, such as silicon [39], alumina [40], silicon nitride [41], SiC–Al2O3

nanocomposites [42] and transition metal nitride coatings [43,44].Generally, the formation of roll-like debris is associated with the tribo-

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 9: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

Fig. 15.Worn surface morphologies and debris of (a, b) as-deposited Co-Ni-P coating and the coatings annealed at (c, d) 400 °C and (e, f) 500 °C.

9C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

chemical reactions of sliding surfaces with the environment species(e.g., water and oxygen) to form a tribofilm [45,46]. Some authors be-lieve that the cylindrical debris on silicon [45] and TiAlN/VN multilayercoatings [43] were generated by the agglomeration and densification of

Table 1Composition of the debris on the worn surface of as-deposited Co-Ni-P coating and thecoatings annealed at 400 °C and 500 °C for 1 h.

Content/at.% Debris

As-deposited sample Annealed samples

400 °C 500 °C

O 37 44 30Co 36 31 39Ni 13 9 10P 14 14 21Fe 0 2 0

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

wear particleswhichwere subjected to opposing tangential forces at thetop and bottom. However, the roll-like debris shown in Fig. 15(b, d, f) isnot solid. Nevertheless, the phenomenon that the plate-like flakes withwarped edges start to detach from the worn surface has been observed.Therefore, it is suggested that the rolls in the present research are madeup of the removed sheets.

It has been accepted that the presence of the rolls is related to themild wear regime, but the effect on the tribological properties is contro-versial. Fisher et al. [40] believed that the rolls on thewear track of a flatAl2O3 specimen against an Al2O3 ball were generally soft, and thereforethey were the consequence rather than the cause of the low wear rate.Bajwa et al. [47] found that the rolls formed by the tribochemical wearhad limited contribution to the wear rate in the wear testing ofSiC-Al2O3 composites against alumina balls. However, Zhou et al.[48] reported that the roll-like wear debris formed during the slidingof TiAlN/VN coatings against Al2O3 balls is associated with higher fric-tion. By contrast, Zanoria et al. [45] observed that the formation of the

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 10: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

10 C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

cylindrical debris on silicon at 600 °C resulted in the decrease of the co-efficient of friction from 0.5 to under 0.2. Moreover, the wear rate wasmuch lower compared to that of the sample tested at 100 °C withoutthe formation of rolls. The authors attributed the reduced coefficientof friction and improved wear resistance to the presence of the rollswhich acted as roller bearings. They also found that in the initial stageof the sliding test, the coefficient of friction was approximately 0.6and then decreased until it reached a steady state (0.1–0.2). The processis similar to the evolution of coefficient of friction of the Co-Ni-P coatingannealed at 500 °C in the present research as shown in Fig. 13. It is sug-gested that the initial fluctuation is due to the high surface roughness ofthe oxide layer as mentioned above. Additionally, the generation ofwear debris and tribofilm leads to the increase of coefficient of frictionin the running-in stage of sliding tests [43]. It is noticed that the roll-like debris on the wear track of the coating annealed at 500 °C havemuch higher population density and more regular shape. The debrisare also aligned more closely perpendicular to the sliding direction.They might be related to the slightly lower coefficient of friction thanthat of the as-deposited sample in the steady-state stage although itwas not an evident contribution. It is an interesting topicwhich requiresmore experimental work on the composition, hardness and the role ofroll-like debris and tribofilm for future work.

4. Conclusion

A new Co-Ni-P coating with reduced coefficient of friction andimproved wear resistance has been developed by electrodepositionfor the replacement of hard chrome coatings. The compositional,morphological, structural, mechanical and tribological properties ofthe as-deposited and annealed Co-Ni-P coatings were characterised.The following conclusions have been drawn:

(1) Along the growth direction, the microstructure changed fromnanocrystalline to amorphous structure as the phosphorous con-tent increased from 7 at.% to 26 at.%. Correspondingly, the sur-face morphology changed from crystal clusters to smoothsurfacewith low surface roughness (Ra = 12 nm). The structuralchange occurred as the P content exceeds approximately15 at.% after plating for 12 min due to the atomic mismatch.The local pH rather than the substrate had a significant effecton the evolution of surface morphologies, composition andmicrostructure.

(2) The oxidation of heat-treated Co-Ni-P coatings started from300 °C, and the top layer of the coating annealed at 500 °Cwas totally covered by the fibrous cobalt oxide, mainly CoO.The layer depleted in Co was found beneath the oxide layer,which indicated that the Co-Ni-P alloys undergo the externaloxidation. Ni12P5 precipitates were formed from the amor-phous phase when the coatings were annealed above 300 °C.The hardness of the as-deposited sample was 602 ± 15 HV.As the heat treatment temperature reached 400 °C, themicrohardness increased to 980 ± 65 HV (comparable tothat of hard chrome coatings) because of the precipitationhardening. Subsequently, it decreases to 650 ± 30 HV afterannealing at 500 °C probably due to the precipitate coarsening,Ni12P5 intermediate phase transition and grain growth.

(3) The coefficients of friction of both the as-deposited Co-Ni-P coat-ing and the heat-treated samples on dry sliding conditions againsta AISI-52100 stainless steel ball were approximately 0.3, only halfof that of hard chromium. As the annealing temperature rose to300 °C, thewear rate started to decrease, and the coating annealedat 400 °C had the lowest wear rate (2.8 × 10−6 mm3N−1 m−1),which was one order of magnitude lower than that of hardchrome coatings. With a further increase of the temperature to500 °C, the wear rate increased, which may be attributed to itsdecreased microhardness and high internal stress.

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

(4) The roll-like debris were found on the worn surface. They werebelieved to be made up of the removed sheets from the coatingsurface. The more roll-like debris with more regular shape wereobserved on the worn surface of the coating annealed at 500 °C.Furthermore, the debris were also aligned more closely per-pendicular to the sliding direction. They may be related tothe slightly lower coefficient of friction compared to that ofthe as-deposited coating in the stable stage. In the run-instage, the coefficient of friction of the coating annealed at500 °C is less stable than that of the as-deposited coating. Itis attributed to its higher surface roughness and the formationprocess of the roll-like debris.

Acknowledgements

The authors would like to thank Ekaterin Minev in University ofCardiff for technical support on FIB and financial support from ICUK.

References

[1] T. Sahraoui, N.E. Fenineche, G. Montavon, C. Coddet, J. Mater. Process. Technol.152 (2004) 43–55.

[2] L.P. Wang, Y. Gao, Q. Xue, H. Liu, T. Xu, Surf. Coat. Technol. 200 (2006) 3719–3726.[3] K. Legg, M. Graham, P. Chang, F. Rastagar, A. Gonzales, B. Sartwell, Surf. Coat.

Technol. 81 (1996) 99–105.[4] K. Legg, in: T.S. Sudarshan, J.J. Stiglich,M. Jeandin (Eds.), ASM International,Materials

Park, Ohio, 2002, pp. 1–10.[5] R.C. Agarwala, V. Agarwala, Sadhana 28 (2003) 475–493.[6] D.H. Jeong, U. Erb, K.T. Aust, G. Palumbo, Scr. Mater. 48 (2003) 1067–1072.[7] B. Bozzini, C. Martini, P.L. Cavallotti, E. Lanzoni, Wear 225 (1999) 806–813.[8] Y. Wu, H. Liu, B. Shen, L. Liu, W. Hu, Tribol. Int. 39 (2006) 553–559.[9] L.P. Wang, Y. Gao, Q. Xue, H. Liu, T. Xu, Appl. Surf. Sci. 242 (2005) 326–332.

[10] C. Ma, S.C. Wang, L.P. Wang, F.C. Walsh, R.J.K. Wood, Wear (2013), http://dx.doi.org/10.1016/j.wear.2013.01.121, (in press).

[11] T.S.N. Sankara Narayanan, S. Selvakumar, A. Stephen, Surf. Coat. Technol. 172 (2003)298–307.

[12] H. Matsuda, G.A. Jones, O. Takano, P.J. Grundy, J. Magn. Magn. Mater. 120 (1993)338–341.

[13] M. Parente, O. Mattos, S. Diaz, P. Neto, F. Miranda, J. Appl. Electrochem. 31 (2001)677–683.

[14] Y. Li, R. Wang, F. Qi, C. Wang, Appl. Surf. Sci. 254 (2008) 4708–4715.[15] J.J. Podesta, R.C.V. Piatti, A.J. Arvia, Int. J. Hydrogen Energy 20 (1995) 111–122.[16] W.L. Liu, W.J. Chen, T.K. Tsai, S.H. Hsieh, S.Y. Chang, Appl. Surf. Sci. 253 (2007)

3843–3848.[17] Y. Huang, K. Shi, Z. Liao, Y. Wang, L. Wang, F. Zhu, Mater. Lett. 61 (2007) 1742–1746.[18] K.S. Lew, M. Raja, S. Thanikaikarasan, T. Kim, Y.D. Kim, T. Mahalingam, Mater. Chem.

Phys. 112 (2008) 249–253.[19] G. Pattanaik, D.M. Kirkwood, X. Xu, G. Zangari, Electrochim. Acta 52 (2007)

2755–2764.[20] D. Kirkwood, V.C. Zoldan, A.A. Pasa, G. Zangari, J. Electrochem. Soc. 157 (2010)

D181–D186.[21] P. Cojocaru, L. Magagnin, E. Gómez, E. Vallés, J. Alloys Compd. 503 (2010) 454–459.[22] S.S. Djokić, J. Electrochem. Soc. 146 (1999) 1824–1828.[23] M. Manna, N. Bandyopadhyay, D. Bhattacharjee, Surf. Coat. Technol. 202 (2008)

3227–3232.[24] T. Mimani, S. Mayanna, N. Munichandraiah, J. Appl. Electrochem. 23 (1993)

339–345.[25] V. Vasilache, G. Gutt, T. Vasilache, Rev. Chim. 59 (2008) 915–920.[26] T.O. Drews, J.C. Ganley, R.C. Alkire, J. Electrochem. Soc. 150 (2003) C325–C334.[27] C.M.A. Brett, A.M.O. Brett, Electrochemistry: Principles, Oxford Science Publications,

New York, Methods and Applications, 1993.[28] A. Bai, C.C. Hu, Electrochim. Acta 50 (2005) 1335–1345.[29] D. Lewis, E.J. Wheeler, D.O. Northwood, J. Educ. Modules, Mater. Sci. Eng. (1982)

673–699.[30] A.W. Burton, K. Ong, T. Rea, I.Y. Chan, Microporous Mesoporous Mater. 117 (2009)

75–90.[31] C. Ma, S.C.Wang, R.J.K.Wood, J. Zekonyte, Q. Luo, F.C.Walsh, Met. Mater. Int. (2013),

(in press).[32] Y. Li, TheOxidation ofNd–Fe–BAlloys. , (Doctoral Thesis) University of Birmingham,

2001.[33] K. Shimizu, K. Kobayashi, P. Skeldon, G.E. Thompson, G.C. Wood, Mater. Sci. Eng. A

Struct. Mater. 198 (1995) 35–41.[34] V. Chapman, B.J. Welch, M. Skyllas-Kazacos, Corros. Sci. 53 (2011) 2815–2825.[35] M.H. Seo, J.S. Kim, W.S. Hwang, D.J. Kim, S.S. Hwang, B.S. Chun, Surf. Coat. Technol.

176 (2004) 135–140.[36] L. Chang, P.W. Kao, C.H. Chen, Scr. Mater. 56 (2007) 713–716.[37] J. Deng, M. Braun, Diam. Relat. Mater. 4 (1995) 936–943.[38] W.M. Rainforth, J. Mater. Sci. 39 (2004) 6705–6721.[39] E. Zanoria, S. Danyluk, M. McNallan, Wear 181 (1995) 784–789.[40] T.E. Fischer, Z. Zhu, H. Kim, D.S. Shin, Wear 245 (2000) 53–60.

x.doi.org/10.1016/j.surfcoat.2013.08.009

Page 11: The electrodeposition and characterisation of low-friction and wear-resistant Co-Ni-P coatings

11C. Ma et al. / Surface & Coatings Technology xxx (2013) xxx–xxx

[41] T.E. Fischer, H. Tomizawa, Wear 105 (1985) 29–45.[42] H.J. Chen, W.M. Rainforth, W.E. Lee, Scr. Mater. 42 (2000) 555–560.[43] Q. Luo, Tribol. Lett. 37 (2010) 529–539.[44] Q. Luo, Wear 271 (2011) 2058–2066.[45] E.S. Zanoria, S. Danyluk, M.J. McNallan, Tribol. Trans. 38 (1995) 721–727.

Please cite this article as: C. Ma, et al., Surf. Coat. Technol. (2013), http://d

[46] V.S.R. Murthy, H. Kobayashi, N. Tamari, S. Tsurekawa, T. Watanabe, K. Kato, Wear257 (2004) 89–96.

[47] S. Bajwa, W.M. Rainforth, W.E. Lee, Wear 259 (2005) 553–561.[48] Z. Zhou, W.M. Rainforth, C.C. Tan, P. Zeng, J.J. Ojeda, M.E. Romero-Gonzalez, P.E.

Hovsepian, Wear 263 (2007) 1328–1334.

x.doi.org/10.1016/j.surfcoat.2013.08.009