surface & coatings technology - veralit - max flexibility. no

7
Homogeneous electroless NiP/SiO 2 nanocomposite coatings with improved wear resistance and modied wear behavior Yoram de Hazan a, , Denise Zimmermann a,b , Markus Z'graggen c , Sigfried Roos d , Christos Aneziris b , Hans Bollier e , Peter Fehr e , Thomas Graule a,b a Laboratory for High Performance Ceramics, Empa, Swiss Federal Laboratories for Materials Testing and Research, Überlandstrasse 129, CH-8600 Dübendorf, Switzerland b Institute of Ceramic, Glass and Construction Materials, Technische Universität Bergakademie Freiberg, Agricolastraße 17, 09599 Freiberg, Germany c IWT, Institut für Werkstofftechnologie, Richtistrasse 15, CH-8304 Wallisellen, Switzerland d Laboratory for Nanoscale Materials Science, Empa, Swiss Federal Laboratories for Materials Testing and Research, Überlandstrasse 129, CH-8600 Dübendorf, Switzerland e Veralit AG, Wagistrasse 7, CH-8952 Schlieren, Switzerland abstract article info Article history: Received 1 March 2010 Accepted in revised form 3 April 2010 Available online 10 April 2010 Keywords: Electroless NiP SiO 2 Nanocomposite coating Hardness Wear resistance Highly homogeneous NiP/SiO 2 nanocomposite coatings were prepared by electroless deposition from composite baths containing well dispersed SiO 2 nanoparticles. The particle concentration in the coatings increases gradually with the SiO 2 concentration in the bath and reaches an observed maximum of 25 vol.% at a bath concentration of 10 g/l. Tribological measurements of the heat treated coatings indicate that while the hardness of the coating is decreased compared to that of the NiP coating without particles, an improvement in Pin-on-Disk wear resistance up to a factor of 2 can be achieved. The aspect ratio of the wear track of samples heat treated above 270 °C is found to increase with the concentration of nanoparticles in the coatings. This observation indicates that the nanoparticles also modify the wear behavior of the coatings. © 2010 Elsevier B.V. All rights reserved. 1. Introduction With the new prospects to improve the properties of composite electro and electroless nickel (EN) coatings by incorporation of submicrometer and nanometer size particles these research elds are receiving increased attention lately [116]. Submicron and nanopar- ticles are able to improve the tribological properties of the coatings through modication of the crystalline growth of the growing nickel lms [15] and creation of dispersion hardening effects at low incorporation levels [1,17]. The potential benets of such reinforcement by nanometer size particles can only be understood and quantied if the particles are sufciently dispersed in the nickel coating. Most ceramic nanoparticle dispersions are unstable in the high ionic strength/high temperature electro and EN plating baths without special surface modication, which is able to prevent particle agglomeration [69]. SiO 2 is one exception to this rule due to its low Hamaker constant in aqueous media implying reduced attraction between particles [18,19] and/or the existence of hydration (solvation) layers which prevent particle agglomeration [1921]. The exploration of nanoparticles to improve the tribological properties of electroless NiP nanocomposite coatings dates back to the work of Metzger et al. [22]. In this comprehensive work which involves NiP/SiC and NiP/Al 2 O 3 also NiP/SiO 2 nanocomposite coatings were prepared and tested for wear resistance. While SiC and Al 2 O 3 have shown improvement of the Taber wear resistance, the incorporation of SiO 2 (in the form of precipitated silica nanopowder) has not yielded an improvement. Fabrication and testing of NiP/SiO 2 nanocomposite coatings is also described in recent work [1015]. Most recently and relevantly, Dong et al. [15] have examined the effect of silane modied SiO 2 on the wear resistance of NiP/SiO 2 coatings. While an improvement of wear resistance with particles was claimed, no comparison of wear resistance with NiP coatings was made for heat treated coatings. Moreover, no conclusion as for the effect of particle concentration in the coatings can be drawn. In this paper we present a systematic investigation on the effect of low cost, commercially available pyrogenic SiO 2 nanoparticles on the tribological properties of high phosphorus NiP/SiO 2 coatings. This work validates our earlier (unpublished) ndings with medium phosphorus NiP/SiO 2 coatings [23]. 2. Materials and methods 2.1. Materials Aerosil OX50 (Evonik Degussa, Germany) a spherical, non aggregated pyrogenic powder with a mean particle size of 40Surface & Coatings Technology 204 (2010) 34643470 Corresponding author. Tel.: +41 (0) 44 823 4817. E-mail address: [email protected] (Y. de Hazan). 0257-8972/$ see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.surfcoat.2010.04.007 Contents lists available at ScienceDirect Surface & Coatings Technology journal homepage: www.elsevier.com/locate/surfcoat

Upload: others

Post on 12-Sep-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Surface & Coatings Technology - Veralit - Max flexibility. No

Surface & Coatings Technology 204 (2010) 3464–3470

Contents lists available at ScienceDirect

Surface & Coatings Technology

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

Homogeneous electroless Ni–P/SiO2 nanocomposite coatings with improved wearresistance and modified wear behavior

Yoram de Hazan a,⁎, Denise Zimmermann a,b, Markus Z'graggen c, Sigfried Roos d, Christos Aneziris b,Hans Bollier e, Peter Fehr e, Thomas Graule a,b

a Laboratory for High Performance Ceramics, Empa, Swiss Federal Laboratories for Materials Testing and Research, Überlandstrasse 129, CH-8600 Dübendorf, Switzerlandb Institute of Ceramic, Glass and Construction Materials, Technische Universität Bergakademie Freiberg, Agricolastraße 17, 09599 Freiberg, Germanyc IWT, Institut für Werkstofftechnologie, Richtistrasse 15, CH-8304 Wallisellen, Switzerlandd Laboratory for Nanoscale Materials Science, Empa, Swiss Federal Laboratories for Materials Testing and Research, Überlandstrasse 129, CH-8600 Dübendorf, Switzerlande Veralit AG, Wagistrasse 7, CH-8952 Schlieren, Switzerland

⁎ Corresponding author. Tel.: +41 (0) 44 823 4817.E-mail address: [email protected] (Y. de Haz

0257-8972/$ – see front matter © 2010 Elsevier B.V. Adoi:10.1016/j.surfcoat.2010.04.007

a b s t r a c t

a r t i c l e i n f o

Article history:Received 1 March 2010Accepted in revised form 3 April 2010Available online 10 April 2010

Keywords:ElectrolessNi–PSiO2

Nanocomposite coatingHardnessWear resistance

Highly homogeneous Ni–P/SiO2 nanocomposite coatings were prepared by electroless deposition fromcomposite baths containing well dispersed SiO2 nanoparticles. The particle concentration in the coatingsincreases gradually with the SiO2 concentration in the bath and reaches an observed maximum of 25 vol.% ata bath concentration of 10 g/l. Tribological measurements of the heat treated coatings indicate that while thehardness of the coating is decreased compared to that of the Ni–P coating without particles, an improvementin Pin-on-Disk wear resistance up to a factor of 2 can be achieved. The aspect ratio of the wear track ofsamples heat treated above 270 °C is found to increase with the concentration of nanoparticles in thecoatings. This observation indicates that the nanoparticles also modify the wear behavior of the coatings.

an).

ll rights reserved.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

With the new prospects to improve the properties of compositeelectro and electroless nickel (EN) coatings by incorporation ofsubmicrometer and nanometer size particles these research fields arereceiving increased attention lately [1–16]. Submicron and nanopar-ticles are able to improve the tribological properties of the coatingsthrough modification of the crystalline growth of the growing nickelfilms [1–5] and creation of dispersion hardening effects at lowincorporation levels [1,17].

The potential benefits of such reinforcement by nanometer sizeparticles can only be understood and quantified if the particles aresufficiently dispersed in the nickel coating. Most ceramic nanoparticledispersions are unstable in the high ionic strength/high temperatureelectroandENplatingbathswithout special surfacemodification,which isable to prevent particle agglomeration [6–9]. SiO2 is one exception to thisrule due to its lowHamaker constant in aqueousmedia implying reducedattraction between particles [18,19] and/or the existence of hydration(solvation) layers which prevent particle agglomeration [19–21].

The exploration of nanoparticles to improve the tribologicalproperties of electroless Ni–P nanocomposite coatings dates back to

the work of Metzger et al. [22]. In this comprehensive work whichinvolves Ni–P/SiC and Ni–P/Al2O3 also Ni–P/SiO2 nanocompositecoatings were prepared and tested for wear resistance. While SiCand Al2O3 have shown improvement of the Taber wear resistance, theincorporation of SiO2 (in the form of precipitated silica nanopowder)has not yielded an improvement. Fabrication and testing of Ni–P/SiO2

nanocomposite coatings is also described in recent work [10–15].Most recently and relevantly, Dong et al. [15] have examined theeffect of silane modified SiO2 on the wear resistance of Ni–P/SiO2

coatings.While an improvement of wear resistance with particles wasclaimed, no comparison of wear resistance with Ni–P coatings wasmade for heat treated coatings. Moreover, no conclusion as for theeffect of particle concentration in the coatings can be drawn.

In this paper we present a systematic investigation on the effect oflow cost, commercially available pyrogenic SiO2 nanoparticles on thetribological properties of high phosphorus Ni–P/SiO2 coatings. Thiswork validates our earlier (unpublished) findings with mediumphosphorus Ni–P/SiO2 coatings [23].

2. Materials and methods

2.1. Materials

Aerosil OX50 (Evonik Degussa, Germany) a spherical, nonaggregated pyrogenic powder with a mean particle size of 40–

Page 2: Surface & Coatings Technology - Veralit - Max flexibility. No

Table 1Pretreatment steps of 100Cr6 steel substrates before Ni–P/SiO2 composite electrolesscoating.

Timemin

Substrate pretreatment procedure

5–10 Ultrasonic degreasing (NaOH)5 Half concentrated HC11 Electiolvtic degreasing solution (NaOH)

(alternate anodic/cathodic)10 V

0.5 Universal activation solution (H2SO4/HCl)0.5 Nickel strike (NiC12/HCl) 5 V2 Electroless NiPlate 500 without particles (87 °C)

3465Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

50 nm and surface area of 50 m2/g was used as the SiO2 nanoparticle[19]. The SiO2 nanoparticles were used as received without anysurface modification. NiPlate500, a commercial high phosphorus ENplating solution was obtained fromMicron s.r.l., Italy. The EN solutionis supplied in two makeup parts (A and B) and an additionalreplenishing part (Part C). Part A contains primarily NiSO4 and parts Band C contain the reducing agent NaH2PO2 along with proprietarycomplexing agents such as citrate and lactate.

2.2. Preparation of dispersions and composite Ni–P bath

20 wt.% colloidal silica dispersions in DI water were prepared byball milling (90 rpm) with 0.5 mm ZrO2 balls for 12 h. Mixing of thecomposite plating bath was performed in a plating vessel, a 5 l glassbeaker, under magnetic stirring in the following order: 1 l DI water,SiO2 dispersion, 300 ml NiPlate500 part A, 450 ml of part B, balance DIwater to 5 l. The initial bath contains 6 g/l nickel and 35 g/l NaH2PO2.The pH of the mixed bath is 4.7±0.1. All mixing procedures are doneat room temperature (RT). The particle size distribution (PSD) wasmeasured with a Beckman Coulter LS230 equipped with polarizationintensity differential scattering (PIDS) for reliable analysis in thesubmicron range. Fig. 1 shows the PSD of the SiO2 particles in DI water(pH=4.5) and after addition to the nickel plating solution. Thedispersion in DI water is monomodal (main peak around 120 nm)indicating a well dispersed system. The absolute particle size given bythe LS230 in the low submicrometer range, however, is not accuratedue to an imperfect optical model [8,9]. An additional peak around300 nm appears in the nickel solution indicating that partialaggregation occurs in the system. The ratio of the two peaks suggeststhat the aggregates may consist of only several primary particles.

2.3. Electroless Ni–P composite plating

Disks with a diameter of 40 mm (0.25 dm2 in area) made of100Cr6 steel were used as substrates for composite EN plating. Oneside of the disk, which was later used for tribological measurementshad a maximum (Rt) and average surface roughness (Ra) of 0.138 μmand 0.012 μm, respectively. Table 1 lists the steps employed in thepretreatment of the 100Cr6 steel substrates. Substrates weredegreased, etched, activated, coated with 0.2 μm nickel strike layerand subsequently coated with Ni–P without particles at 87 °C for2 min directly before composite plating. This procedure was foundnecessary to obtain good interface for composite coating. The mixedcomposite bath was heated to plating temperatures of 86–87 °C. 5pretreated samples having a total surface of 1.25 dm2 hanged on asteel rod were placed in the bath immediately after the pretreatment.

Fig. 1. SiO2 particle size distribution in DI water (25% dispersion, pH=4.5) and in nickelplating solution (5% dispersion in Niplate500), measured at RT.

The bath was mixed by a magnetic stirrer (100–200 rpm) and kept attemperature (±1 °C) with a temperature controlled heating jacket.The pH was kept at 4.7±0.1 by frequent additions of 25% ammoniasolutions. No special measures to prevent particle agglomeration(such an in situ ultrasound [16]) have been employed or foundnecessary during codeposition.

2.4. Characterization of nanocomposite coatings

Particle size distribution and concentration in the Ni–P/SiO2

composite coatings were evaluated by SEM/EDS (LEO 1455/Oxfordinstruments). For the examinations, cross sections of the disks wereembedded in a conductive resin. The metallographic preparationfollowed a standard procedure. The last polishing step was 1 µmDiamond. The SEM conditions for all measurements were: a workingdistance of 15–20 mm, an acceleration voltage of 20 kV and a probecurrent of 350–500 pA. The pictures were taken using a BSE Detector.The EDS signal was calibrated with a cobalt standard before eachseries of measurements. Each experimental point is the average of asection of thickness of 2 µm and a length of 100 µm. The tribologicalproperties of the coatings were measured after three heat treatmentprocedures conducted in air: a) 270 °C for 10 h, b) 390 °C for 2 h andc) 480 °C for 1 h. The samples treated at 390 °C or 480 °C in air possessa thin (100–200 nm) oxide layer at their surface. This thin oxide layerhas not been removed prior to hardness and PODmeasurements sinceits effect on the results was negligible and its removal would havemodified the surface roughness.

Vickers hardness (Melit 430SVD, Switzerland) was measured at aforce of 3 N (HV0.3) for 15 s. The hardness was measured for a set of 9points across the diameter of the disks from which a mean value andstandard deviation were calculated.

The wear resistance of Ni–P composite coated disks was measuredagainst a 10 mm 100Cr6 steel ball by the Pin-on-Disk (POD) method(CSEM Tribometer, Switzerland). Sample rotation speed at a radius of11 mm was 20 cm/s. A load of 10 N was used. Experiments with totaldistance of 11.5 km were made (16 h). Such long distance wasnecessary to obtain statistically valid data (weight loss in the range of0.5 mg ±0.1 mg). No lubricant was used. Ambient temperature waskept at 25 °C and humidity at 50% in all experiments. The weight lossof the samples and the ball during the test was measured. The weightchange of the ball was very small in all experiments, within themeasurement error and therefore not reported here.

The volumetric wear rate of the samples was obtained using theSiO2 particle concentration measured by EDS, assuming the density ofthe coating obeys the rule of mixtures: 1

ρcomposite=

XSiO2ρSiO2

+1−XSiO2

ρNiPwhere

XSiO2is the weight fraction of SiO2 in the Ni–P/SiO2 nanocomposite

coating and ρSiO2and ρNi–P are taken as 2.2 and 8.0 g/cm3, respectively.

A universal wear rate factor (WRF) can be calculated by thefollowing equation:

WRF =Vremoved

F⋅lweartrack;

m3

N⋅m

" #ð1Þ

Page 3: Surface & Coatings Technology - Veralit - Max flexibility. No

3466 Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

where Vremoved is the volume of substrate removed during the PODtest, F is the force applied and lweartrack is the distance traveled.

Tencor P-10 surface profiler (Tencor, USA) was used to analyze thePOD wear track (depth, width and shape) of the tested samples. Thewear track dimensions were measured at four equally spacedlocations from which a mean value and a standard deviation werecalculated. The volume of the wear track was also estimated bymeasuring the race track cross sectional area andmultiplying it by thewear track length. The wear track area was measured by printing theprofiles on a paper, cutting and weighing the papers. The weightswere normalized with a weight of paper with a known area.

3. Results and discussion

3.1. Preparation of homogeneous Ni–P/SiO2 nanocomposite coatings

Fig. 2 shows the SEM cross section of Ni–P/SiO2 nanocompositecoatings plated for 3 h from a high phosphorus bath containing 1–20 g/l SiO2 nanoparticles (the 100Cr6 steel substrate is visible at thebottom of the SEM images). All coatings produced are between 25 and30 μm thick and exhibit homogeneous distribution of nanoparticles intheir cross section.

The coatings viewed at higher resolution (Fig. 3) show that thenanoparticles are well dispersed in the coatings at all SiO2 bathconcentrations studied. Most large agglomerates seen in the coatingsare in the range of 100–150 nm, consisting of only several primaryparticles. This is in qualitative agreement with the particle sizedistribution shown in Fig. 1. It is interesting to note that coatingsdeposited from baths with lower SiO2 concentrations (1 and 3 g/l)exhibit somewhat larger aggregates.

Fig. 2. Cross section of Ni–P/SiO2 composite coatings (100Cr6 steel substrate is at the bot

Fig. 3 also shows a gradual increase in particle concentration in theNi–P/SiO2 nanocomposite coatings with particle concentration in thebath. The average SiO2 particle concentration in the coatings,measured by EDS analysis is shown in Fig. 4 as a function of theSiO2 concentration in the plating bath. The cross sectional area usedfor averaging was 100 μm×25 μm. The concentration of SiO2 particlesin the coating increases gradually and predictably to a maximum levelof 25 vol.%, observed experimentally at 10 g/l. At 20 g/l a smalldecrease to 22 vol.% is seen. The experiments at 1 and 10 g/l have beenrepeated with identical findings, confirming the high incorporationlevels of 25 vol.%. At 20 g/l SiO2 concentration in the plating bathmacroscopic agglomerates of SiO2, forming presumably at the liquidbath/air interface are observed. At the same time, an interpolationbetween 3 and 10 g/l SiO2 in Fig. 4 suggests that coatingswith 22 vol.%can more conveniently be produced from baths with a SiO2

concentration of ∼7 g/l.

3.2. Wear behavior of Ni–P/SiO2 nanocomposite coatings

The Vickers hardness values (HV0.3) of the Ni–P/SiO2 coatings afterheat treatment procedures of 270 °C/10 h, 390 °C/2 h and 480 °C/1 hare shown in Fig. 5 as a function of particle concentration in theplating bath. Due to the relatively high standard error only someconclusions can be made. Counter to other findings (e.g. reference[15]), the Ni–P coatings without particles exhibit the highest hardnessvalues after all heat treatments. It is not clear if this discrepancy isrelated to differences in the plating solution, particle surfacemodification or perhaps the force used during HV measurements(3 N vs. 0.5 N). A clear dependence of hardness on the particleconcentration in the bath is apparent at 480 °C. Interestingly, the

tom). a) 1 g/l SiO2, b) 3 g/l SiO2, c) 10 g/l SiO2 and d) 20 g/l SiO2. Scale bar is 20 μm.

Page 4: Surface & Coatings Technology - Veralit - Max flexibility. No

Fig. 3. Cross section of Ni–P/SiO2 composite coatings shown in Fig. 2 viewed at higher resolution. a) 1 g/l SiO2, b) 3 g/l SiO2, c) 10 g/l SiO2 and d) 20 g/l SiO2. Scale bar is 2 μm.

3467Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

maximum hardness which is normally observed after heat treatmentat 400 °C [15], is not observed here for the Ni–P coating withoutparticles after heat treatment at 390 °C.

Due to poor adhesion to steel, the as deposited coatings on 100Cr6steel are unstable without heat treatment above ∼200 °C. The asdeposited coatings are destroyed in a matter of minutes during PODtests. Fig. 6 shows the POD wear resistance of the composite coatingsas a function of particle concentration in the bath and heat treatmenttemperature. The gravimetric wear rate based on sample weight lossduring the POD test (Fig. 6a) is converted to volume loss bynormalizing to the density of the samples (See Section 2.4) andpresented as the volumetric wear (Fig. 6b). Despite this correction,

Fig. 4. Effect of SiO2 particle concentration in the plating bath on SiO2 incorporation inthe Ni–P/SiO2 nanocomposite coating.

Fig. 6a and b show qualitatively the same picture. The Ni–P/SiO2

nanocomposite coatings show in general superior wear resistancecompared to Ni–P coating without particles at all heat treatmenttemperatures studied (the only two exceptions in the data are 20 g/lat 270 °C and 1 g/l at 480 °C).

Table 2 lists the best improvement in wear resistance by consideringthe ratio of the wear rate factorsWRFNi–P/WRFcomposite. An improvementby a factor of 1.5 is found at 270 °C (at 15 vol.% nanoparticles) and 2 at390 °C (21 vol.% nanoparticles). The best improvement appears to bealways at or around15 vol.%nanoparticles in the coatings (corresponding

Fig. 5. Effect of SiO2 particle concentration in the plating bath and heat treatmenttemperature on hardness of the Ni–P/SiO2 nanocomposite coatings.

Page 5: Surface & Coatings Technology - Veralit - Max flexibility. No

Fig. 6. Effect of SiO2 particle concentration in the plating bath and heat treatmenttemperature on the wear rate of the Ni–P/SiO2 nanocomposite coatings. a) Gravimetric,and b) volumetric.

3468 Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

to bath concentration of 3 g/l), irrespective of the heat treatmenttemperature.

After heat treatment at 270 °C, the coating from bath containing3 g/l SiO2 represents a minimum in the wear rate curve (Fig. 6)whereas at 390 °C it represents the beginning of a plateau. In any case,coatings made from baths with 1 g/l SiO2 or less show consistentlyinferior wear resistance.

Using the universal wear rate factor, one can compare absoluteresults from different systems (although not without great caution).In any case, the WRF found here are around 10−15 for the Ni–P and 510−16 for the nanocomposites over 11.6 km. For comparison, Donget al. [15] show the best value of WRF of 10−14, 10–20 times higherthan in the present work. The exact measurement distance is notreported but may be as short as 12 m judging from the data presentedin Fig. 5 of reference [15].

The friction coefficient (average of last 4 km) as a function ofparticle concentration in the bath and heat treatment temperature is

Table 2Largest improvement in wear resistance relative to Ni–P coatings after different heattreatment procedures.

Thermal treatment SiO2 in coatings(vol.%)

WRF(m3/Nm)

WRFNiP/WRFcomposite

270 °C, 10 h 0 1.1E-15 1.515 7.3E-16

390 °C, 2 h 0 1.SE-15 221 7.7E-16

480 °C, 1 h 0 8.7E-16 1.815 4.9E-16

shown in Fig. 7. After heat treatment at 270 °C the coatings show afriction coefficient of ∼0.4 independent of particle concentration. Heattreatment at higher temperatures results in an increase of this value.The friction coefficients of the Ni–P/SiO2 nanocomposite are alwayslower than those of the Ni–P coatings without exception. A sharpincrease of the friction coefficient is seen after heat treatment at390 °C by 49% (0.61) for the Ni–P without particles and 35% and 26%for the coatings produced from baths containing 10 and 20 g/l SiO2,respectively. This change, which may be attributed to increasedroughness after crystallization of Ni3P at 390–400 °C [24] is much lesssignificant for coatings having lower SiO2 concentration (13% and 10%for 1 and 3 g/l, respectively). After heat treatment at 480 °C thefriction coefficient of the nanocomposites show only little variationwith particle concentration.

The lower friction coefficients of the Ni/SiO2 nanocomposites arelikely to contribute to a reduction in wear rate compared to Ni–Pcoatings. At the same time, the trends seen in Fig. 6 are quite differentthan those seen in Fig. 7, therefore excluding friction coefficient frombeing the dominant factor determining wear resistance. For exampleone can observe that the 1 and 3 g/l coatings have similar frictioncoefficients (Fig. 7) but very different wear resistance (Fig. 6).Compared to the friction coefficient measured here for the compositecoatings, which are in the 0.4–0.45 range, Dong et al. [15] find muchlower friction coefficients in the range of 0.23–0.3. As stated before,this may be due to the very different POD measurement distance andtherefore thickness of coating probed.

One aspect of the wear behavior which is not frequently reportedis the wear track shape and dimensions. Fig. 8 presents the width anddepth of the wear track as a function of particle concentration in thebath after different heat treatment temperatures. Similar to thefriction coefficient (Fig. 7), the dimensions of the wear track areindependent of particle concentration in the coating after heattreatment at 270 °C (Fig. 8a). Also here the highest sensitivity of thewear track dimensions to particle concentration is seen after heattreatment at 390 °C (Fig. 8b). The general trend is an increase in depthand a decrease in width of the wear track with increased particleconcentration in the coating. This trend continues after heattreatment at 480 °C, but the variation with particle concentration ismuch diminished (Fig. 8c). Fig. 9 illustrates two extreme examples ofwear track profiles, both obtained after heat treatment at 390 °C.While the 1 g/l SiO2 sample exhibits a relatively shallow and broadwear track, the 10 g/l SiO2 shows a deep and narrow track (note thedifferent scales in the x and y axes). It is important to note that thewidth and depth of the race track at 270 °C do not predict the largereduction in volume loss seen with the nanoparticles (Fig. 6 and

Fig. 7. Effect of SiO2 particle concentration in the plating bath and heat treatmentprocedure on the friction coefficient of the Ni–P/SiO2 nanocomposite coatings.

Page 6: Surface & Coatings Technology - Veralit - Max flexibility. No

Fig. 8. Effect of SiO2 particle concentration in the plating bath on the depth and width ofthe wear track of the Ni–P/SiO2 nanocomposite coatings. (a) Heat treatment at 270 °C,(b) heat treatment at 390 °C, and (c) heat treatment at 480 °C.

Fig. 9. Comparison of wear track profiles of Ni–P/SiO2 nanocomposite coatings heattreated at 390 °C. Top — 7.5 vol.% SiO2 (1 g/l), and bottom — 25 vol.% SiO2 (10 g/l).

Fig. 10. Volume of wear track after heat treatment at 270 °C estimated fromweight lossduring POD tests or by integration of wear track cross sectional area.

3469Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

Table 2). A more precise estimation of the volume loss during PODtests involves the estimation of wear track cross sectional area. Fig. 10shows the volume loss during POD test of samples heat treated at270 °C, estimated from POD weight loss data (from Fig. 6b) or byestimation of volume from the profiles of the wear track. Reasonableagreement is found in Fig. 10, predicting a somewhat lowerimprovement of wear resistance by a factor of 1.35 compared to 1.5shown in Table 2. Good agreement is also found after heat treatmentat 390 °C and 480 °C.

The temperature dependency seen in Fig. 8 is in general quitesimilar to the dependence of friction coefficient on particle concen-tration (Fig. 7). In other words, the dimensions of the wear track andthe friction coefficient may be influenced by material properties.Alternatively, the dimensions of the race track may determine thefriction coefficient.

Page 7: Surface & Coatings Technology - Veralit - Max flexibility. No

Fig. 11. Effect of SiO2 particle concentration in the Ni–P/SiO2 nanocomposite coatingsand heat treatment procedure on the aspect ratio of the wear track.

3470 Y. de Hazan et al. / Surface & Coatings Technology 204 (2010) 3464–3470

The information shown in Fig. 8 can also be represented by a singleparameter, the aspect ratio of the wear track which is defined as the(depth of wear track)/(width of wear track). The variation of theaspect ratio of the wear track (AR) with the SiO2 particle concentra-tion in the Ni–P/SiO2 coating is plotted in Fig. 11. The AR isindependent on SiO2 concentration after heat treatment at 270 °C.However, after heat treatment at higher temperatures (mostdramatically at 390 °C) an increase of the AR with particle concen-tration in the coating is observed above 7.5 vol.% SiO2. The absolutevalue of the AR, ranges between 0.01 and 0.14. This systematic change,which to our knowledge is observed here for the first time, indicates avariation of the wear behavior of the Ni–P/SiO2 nanocompositecoatings with embedded nanoparticles. This perhaps is indicative of agradual shift to the adhesive wear regime with increase in SiO2

nanoparticle concentration in the coating and may dictate the specificapplications of the coatings. The Ni–P coatings without particles showa steady increase in the aspect ratio with heat treatment temperature.Interestingly, at 7.5 vol.% SiO2 (1 g/l) the aspect ratio is independent ofthe heat treatment temperatures studied. The analysis above showsthat nanocomposite systems have complex wear behavior and shouldnot be characterized by volumetric wear rate alone. Although the Ni–P/SiO2 nanocomposite coatings exhibit significantly lower wear rates,due to the difference in their aspect ratios, the depth of the wear trackmay not be reduced in a similar way. In fact, an increase in the depthof the wear track with the incorporation of SiO2 nanoparticles is alsoobserved in Fig. 8. It is important to note that while POD analysisprovides a good guide for wear behavior, specific tests are necessaryfor specific applications. For these, the benefit of the Ni–P/SiO2

nanocomposite coatings would need to be evaluated independently.

4. Conclusions

Homogeneous Ni–P/SiO2 nanocomposites were prepared fromstable composite electroless baths containing SiO2 nanoparticles. The

particle concentration in the coatings increases gradually with theSiO2 concentration in the bath and reaches a maximum of 25 vol.% at10 g/l. The hardness of the heat treated coating is decreased comparedto Ni–P without particles. At the same time an improvement in Pin-on-Disk wear resistance by factors of 1.5 and 2 is achieved after heattreatment at 270 °C and 390 °C, respectively. The friction coefficientwhich is independent of particle concentration after heat treatment at270 °C can also be markedly improved by the nanoparticles after heattreatment at higher temperatures. The increase of the aspect ratio ofthe wear track with concentration of nanoparticles in the Ni–P/SiO2

coatings indicates that the nanoparticles also influence the wearbehavior of the coatings. The benefit of this different wear behaviorneeds to be assessed independently for specific applications.

Acknowledgments

This work was supported by the Swiss KTI Project No. 8511.2.

References

[1] C.T.J. Low, R.G.A. Wills, F.C. Walsh, Surf. Coat. Technol. 201 (2006) 371.[2] F. Erler, C. Jacob, H. Romanus, L. Spiess, B. Wielage, T. Lampke, S. Steinhäuser,

Electrochim. Acta 48 (2003) 3063.[3] J.N. Balaraju, K.S. Kalavati, Rajam, Surf. Coat. Technol. 200 (2006) 3933.[4] Gao Jiaqiang, Wu. Yating, Liu Lei, Hu. Wenbin, Mater. Lett. 59 (2005) 391.[5] C. Gheorghies, G. Carac, I.V. Stasi, J. Optoelectron. Adv. Mater. 8 (2006) 1234.[6] Li. Chen, Liping Wang, Zhixiang Zeng, Junyan Zhang, Mater. Sci. Eng. 434 (2006)

319.[7] Sheng-Lung Kuo, Yann-Cheng Chen, Ming-Der Ger, Wen-Hwa Hwu, Mater. Chem.

Phys. 86 (2004) 5.[8] Yoram de Hazan, Torben Reuter, Dennis Werner, Rolf Clasen, Thomas Graule, J.

Coll. Interf. Sci. 323 (2008) 293.[9] Yoram de Hazan, Dennis Werner, Markus Z'graggen, Michael Groteklaes, Thomas

Graule, J. Colloid Interface Sci. 328 (2008) 103.[10] Ramesh Chandra Agarwala, Vijaya Agarwala, Rahul Sharma, Metal-Org. Nano-

Metal Chem. 36 (2006) 493.[11] Xianghua Yu, Hongyan Wang, Zhenrong Yang, Ping Yin, Xinquan Xin, Appl. Surf.

Sci. 158 (2000) 335.[12] M. Oberseider, C. Jakob, M. Petrova, Z. Noncheva, J. Schawohl, Galvanotechnik 96

(2005) 1214.[13] M. Petrova, Z. Noncheva, E. Dobreva, 51st Internationales Wissenschaftliches

Kolloquium Technische Universität Ilmenau, September 11–15, 2006.[14] Zhang Hui, Liu Rongli, Sen'i Gakkaishi 64 (2008) 372.[15] D. Dong, X.H. Chen, W.T. Xiao, G.B. Yang, P.Y. Zhang, Appl. Surf. Sci. 255 (2009)

7051.[16] K. Dünkel, D. Meyer, T. Lampke, S. Steinhäuser, Materialwissenschaft und

Werkstofftechnik 40 (2009) 888.[17] J.R. Roos, J.P. Celis, H. Kelchtermans, Thin Solid Films 54 (1978) 173.[18] Lennart Bergström, Adv. Coll. Interf. Sci. 70 (1997) 125.[19] Maciej Wozniak, Thomas Graule, Yoram de Hazan, Dariusz Kata, Jerzy Lis, J. Eur.

Ceramic Soc. 29 (2009) 2259.[20] V.I. Zarko, A.A. Tumanov, Reaction Kin. Catal. Lett. 50 (1993) 189.[21] S. Song, C. Peng, M.A. Gonzalez-Olivares, A. Lopez-Valdivieso, T. Fort, J. Coll. Interf.

Sci. 287 (2005) 114.[22] Willy Metzger, Rudi Ott, Gunter Pappe, Helmut Schmidt, US Patent 3753667

(1973).[23] Thomas Graule, Yoram de Hazan, Peter Fehr, Hans Bollier, European Parent

Application EP09005918 (filed April 29, 2009), 2009.[24] J.N. Balaraju, T.S.N. Sankara Narayanan, S.K. Seshadri, J. Appl. Electrochem. 33

(2003) 807.