You are on page 1of 7

Surface & Coatings Technology 204 (2010) 3804–3810

Contents lists available at ScienceDirect

Surface & Coatings Technology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / s u r f c o a t

Ni–TiO2 nanocomposite coating with high resistance to corrosion and wear


P. Baghery ⁎, M. Farzam, A.B. Mousavi, M. Hosseini
Technical Inspection Engineering Department, Petroleum University of Technology, Abadan, Iran

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

Article history: Ni–TiO2 nanocomposite coatings with various contents of TiO2 nanoparticles were prepared by electrodeposi-
Received 26 March 2010 tion in a Ni plating bath containing TiO2 nanoparticles to be codeposited. The influences of the TiO2 nanoparticle
Accepted in revised form 26 April 2010 concentration in the plating bath, the current density and the stirring rate on the composition of nanocomposite
Available online 2 May 2010
coatings were investigated. The composition of coatings was studied by using energy dispersive X-ray system
(EDX). The wear behavior of the pure Ni and Ni–TiO2 nanocomposite coatings were evaluated by a pin-on-disc
Keywords:
Wear
tribometer. The corrosion performance of coatings in 0.5 M NaCl, 1 M NaOH and 1 M HNO3 as corrosive
Corrosion solutions was investigated by potentiodynamic polarization and electrochemical impedance spectroscopy
Nanocomposite coating methods (EIS). The microhardness and wear resistance of the nanocomposite coatings increase with increasing
Potentiodynamic polarization of TiO2 nanoparticle content in the coating. With increasing of TiO2 nanoparticle content in the coating, the
Pin-on-disc tribometer polarization resistance increases, the corrosion current decreases and the corrosion potential shifts to more
EIS positive values.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction site coatings disappear. Therefore, dispersion of these particles in the


plating bath is an important parameter which is affected by nano-
The surface is the most important part of any engineering particle concentration in the plating bath and stirring rate [14,15].
component. It is well known that most components fail from surface In this paper, the electrodeposition technique has been applied
initiated defects such as wear, corrosion, fatigue or fracture [1]. using TiO2 nanoparticles in Ni plating bath for the production of the
Applying a coating on the surface is a method to improve the Ni–TiO2 nanocomposite coatings. The effects of stirring rate, current
surface properties. Several techniques have been applied to produce density and TiO2 nanoparticle concentration in the bath on the
protective coatings, such as electroplating, plasma thermal spray and composition of the deposits are examined. The goal is to produce hard
physical vapor deposition methods [2]. Electroplating composite Ni–TiO2 nanocomposite coating with high resistance to corrosion and
coating is an effective method to prepare composite coating through wear.
the codeposition of metallic, nonmetallic or polymer particles with
metal to improve properties such as corrosion resistance, hardness 2. Materials and methods
and wear performance [3,4]. The size of codeposited particles has an
important effect on the coating properties. The coating properties Ni–TiO2 nanocomposite coatings were deposited on low carbon steel
improved with decreasing of codeposited particles size. Recently, with substrate by the conventional direct current electrodeposition process.
the development of nanotechnology, a new type of composite coating The electroplating bath compositions are as follows: 260 g/l NiSO4·6H2O,
called nanocomposite coating has been developed. These coatings 60 g/l NiCl2·6H2O, 32 g/l H3BO4 and certain content of titanium dioxide
have been applied by electro-codeposition of inert nano-scale nanoparticles (Degussa P-25 anastase) with a particle size of 25 nm. The
particles with metal [5–12]. The codeposited nanoparticles improve substrates were ultrasonically cleaned in acetone for 10 min, dipped in
the coatings performance. acid (10% HCl) and finally washed with distilled water before
The coating properties have been affected by electroplating electrodeposition process. During the codeposition process the bath
conditions such as the plating bath composition, current density and was stirred by a magnetic stirrer in order to keep the particles dispersed
stirring rate [13,14]. Nanoparticles agglomerated very easily in the and prevent sedimentation in the electrolyte suspension. The bath
plating bath which results in enhanced amount of agglomerated temperature was maintained at 50 °C by an automatic controller. The
particles in composite coatings and unique properties of nanocompo- anode was a Pt plate with the size of 20 × 20 mm. The surface mor-
phology and the composition of coatings were characterized with
scanning electron microscopy (SEM) with energy dispersive analyzer
⁎ Corresponding author. Tel.: +98 918 311 7959; fax: +98 631 442 3520. system (EDX). The weight fraction of TiO2 was determined by the
E-mail address: pouria.baghery@yahoo.com (P. Baghery). chemical formula.

0257-8972/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2010.04.061
P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810 3805

2.1. Test methods

2.1.1. Microhardness measurements


The Vickers microhardness measurements were carried out with
loads of 30 g and indentation time of 30 s. The corresponding final
values were determined as the average of 10 measurements.

2.1.2. Wear test


The wear tests were performed on a pin-on-disc type tribometer
with a constant rotation speed of 75 rpm at a constant radius of 1 cm
and a load of 50 N at room temperature. The pin was a bearing steel
(AISI 52100) with a radius of 2.5 mm and a pin tip air radius of 10 mm.
The disc was the low carbon steel coated with a nanocomposite film.
The wear tests were performed under non-lubricated condition with
the total distance of 60 m. The friction loads vs. time were measured
automatically by a loadcell during the test with 5 measurements per
second. The samples were ultrasonically washed in acetone before
and after each test. The weight loss of the samples, to an accuracy of
0.1 mg, was measured and the final value reported was the average of
three test measurements so as to minimize data scattering.

2.1.3. Potentiodynamic polarization test


Potentiodynamic polarization measurements were carried out in an
open to air conventional three electrode cell, containing 500 ml of
electrolyte. Measurements were performed in 0.5 M NaCl and 1 M NaOH
solutions at the temperature of 25 °C. Coated specimens were used as
working electrode and platinum electrode and saturated Ag/AgCl
electrode were used as counter electrode and the reference electrode.
Polarization studies were conducted using computer controlled Autolab
Fig. 1. SEM images of (a) Ni coating (b) Ni–8.3 wt.% TiO2.
PGSTAT 302N at a scan rate of 0.5 mV/s. Autolab GPES (general purpose
electrochemical system) software was used for evaluating the experi-
mental data. The open circuit potential (OCP) was measured after surface (Fig.1a). The Ni–TiO2 nanocomposite coating shows smother
immersion and when OCP reaches to stable condition the polarization surface, more uniform and compact in appearance compare to that of
measurements were done. The corrosion potential (Ecorr) and corrosion pure Ni coating, which indicates that the codeposited TiO2 nanopar-
current density (icorr) were calculated from the intersection of the ticles were uniformly distributed in the Ni matrix of nanocomposite
cathodic and anodic Tafel curves using the Tafel extrapolation method. coating.
The polarization resistance (Rp) was determined using Stren–Geary
equation [16]: 3.1. Electrodeposition factors affecting the composition of coating

βa βc The variation, of weight percentage of codeposited TiO2 with the


Rp = ð1Þ
2:303icorr ðβa + βc Þ concentration of TiO2 nanoparticles in the bath at a current density
of 5 A/dm2 and a stirring rate of 180 rpm, is shown in Fig. 2. It is
where βa and βc are the anodic and cathodic Tafel slopes. The corrosion indicated that the wt.% of TiO2 nanoparticles embedded in coating is
rate (rcorr in millimeter per year) was calculated through the following dependent on the TiO2 nanoparticle concentration in the bath. The
equation [17]: wt.% of TiO2 in coating increases with increasing of TiO2 nanoparticle
concentration in the electrolyte up to 30 g/l. But beyond this value,
icorr M the wt.% of TiO2 in coating declines with increasing of TiO2
rcorr = 0:00327 ð2Þ
nd nanoparticle concentration in the electrolyte. The increasing of

where M, n, and d are molar mass, charge number, and density of tested
metal, respectively.

2.1.4. Electrochemical impedance spectroscopy (EIS)


Electrochemical impedance spectroscopy (EIS) measurements
were performed between 10 mHz and 100 kHz frequency range
using computer controlled Autolab PGSTAT 302N at the temperature
of 25 °C. Autolab FRA (frequency reference analyzer) software was
used for evaluating the experimental data and calculation of
polarization resistance (Rp). The peak-to-peak amplitude of the
sinusoidal voltage signal was 10 mV. Measurements were performed
after 1 h immersion to reach steady state condition.

3. Results and discussions

SEM images of the pure Ni and Ni–TiO2 nanocomposite coating are Fig. 2. Wt.% of codeposited TiO2 vs. the TiO2 concentration in the bath (current
indicated in Fig. 1. The pure Ni coating demonstrated a rather regular density = 5 A/dm2, T = 50 °C, stirring rate = 180 rpm).
3806 P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810

Fig. 3. Wt.% of codeposited TiO2 vs. the current density (TiO2 concentration in Fig. 5. Variation of microhardness with wt.% of codeposited TiO2.
bath = 30 g/l, T = 50 °C, stirring rate = 180 rpm).

TiO2 wt.% in coating at the first step can be explained by Guglielmi's strongly affects the wt.% of TiO2 nanoparticles since nanoparticles
two-step adsorption model [18,19]. The inert nanoparticles, once should be transported to the cathode surface for codeposition. TiO2
adsorbing metal ions, acquire positive surface charge and are nanoparticle content in coating increase with stirring rate and reach a
potentially adsorbed on the electrode surface. Further reduction of maximum value at 180 rpm, then decreases with increasing the
these metal ions results in strong adsorption of the particles onto the stirring rate. At stirring rate lower than 180 rpm, the fluid flow is not
metal matrix [20]. The agitation of electrolyte has a significant role in capable of transporting all the nanoparticles to the cathode surface
the codeposition of the inert nanoparticles in metal electrodeposi- and the codeposition behavior of TiO2 nanoparticles is apparently
tion as same as the ion adsorption [21,22]. In other words, the ion controlled by nanoparticle transfer. When the stirring rate is too high,
adsorption rate increases with increasing of TiO2 nanoparticle the decreasing trend of the weight percentage is principally caused by
concentration in electrolyte, thus resulting in a higher wt.% of TiO2 the collision factor [24].
content in coating. The decreasing of TiO2 wt.% in coating at the
second step is attributed to the agglomeration of TiO2 nanoparticles
in the electrolyte owing to their poor wettability [23].
Fig. 3 shows the relationship between the current density and the
content of codeposited TiO2 nanoparticles. The wt.% of TiO2 nanopar-
ticles in coating is enhanced by increasing of current density up to 5 A/
dm2 and beyond this current density, the TiO2 nanoparticle content in
coating decreased. The increasing of TiO2 nanoparticles wt.% in coating
at a current density less than 5 A/dm2, is due to the increasing of the
adsorbed nanoparticles tendency to arrive in the cathode surface. The
process is controlled by the adsorption of nanoparticles. At a current
density higher than 5 A/dm2, the metal ions movement to the cathode
surface is faster than adsorbed nanoparticles and on other hand, more
rapid reduction of metal ions to metal matrix atoms on the cathode
surface decrease TiO2 nanoparticle content in coating.
The relationship between the stirring rate and the content of
codeposited TiO2 nanoparticles is shown in Fig. 4. The stirring rate
Fig. 6. Friction coefficient curves of (a) pure Ni coating, (b) Ni–6.5 wt.% TiO2 and
(c) Ni–8.3 wt.% TiO2.

Fig. 4. Wt.% of codeposited TiO2 vs. the stirring rate (TiO2 concentration in bath = 30 g/l,
current density = 5 A/dm2, T = 50 °C). Fig. 7. Variation of wear loss with wt.% of codeposited TiO2.
P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810 3807

Fig. 8. Tafel polarization curves for pure Ni coating and Ni–TiO2 nanocomposite coatings in 1 M NaCl solution.

Fig. 9. Tafel polarization curves for pure Ni coating and Ni–TiO2 nanocomposite coatings in 1 M NaOH solution.

3.2. Hardness measurements coatings. The mechanisms of strengthening of coatings are the grain
refinement strengthening and the dispersion strengthening [26,27].
The microhardness of Ni–TiO2 nanocomposite coatings and pure
Ni coating under same condition are indicated in Fig. 5. It was found 3.3. Wear behavior of coatings
that the hardness values of Ni–TiO2 nanocomposite coatings are
higher than that of pure Ni plating (248 Hv) prepared under the same Fig. 6 shows the friction coefficient of pure Ni coating and Ni–TiO2
condition and for a rising content of embedded TiO2 nanoparticles in nanocomposite coatings containing different TiO2 wt.%, under the
coating, an increase in the microhardness was observed, similar to identical test conditions. The friction coefficients of all coatings are
what has been reported on Ni–Si3N4 [7] and Ni–WC [25] composite approximately same at the first 10 m. Thereby, the pure Ni coating

Fig. 10. Tafel polarization curves for pure Ni coating and Ni–TiO2 nanocomposite coatings in 1 M HNO3 solution.
3808 P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810

friction coefficient increases dramatically to 1 and then maintains at a


constant level. However the friction coefficients of Ni–TiO2 nano-
composite coatings exhibited little change and keep stable during the
test. The hard nano-scale reinforcements embedded in nanocompo-
site coatings reduce direct contact between metal matrix and abrasive
surface. On other hand, the nanoparticles, separated from matrix due
to abrasion, act as solid lubricants between two wear surfaces. Fig. 7
shows the wear loss of pure Ni coating and Ni–TiO2 nanocomposite
coatings with different TiO2 contents. It is obvious that wear
resistance of nanocomposite coatings are more than pure Ni coating.
This can be attributed to the strengthening effect and friction
coefficient decreasing which was explained previously. The hardness
and friction coefficient are two parameters which affect the wear
resistance [28].

3.4. Potentiodynamic polarization Fig. 11. Nyquist impedance diagrams for pure Ni coating and Ni–TiO2 nanocomposite
coatings in 1 M NaCl solution.
The potentiodynamic polarization curves of pure Ni and Ni–TiO2
nanocomposite coatings in 0.5 M NaCl, 1 M NaOH and 1 M HNO3
solutions are presented in Figs. 8–10. Corrosion characteristics such as 3.4.1. Passive layer formation in HNO3 solution
corrosion potential (Ecorr), corrosion current (icorr) and anodic/cathodic Fig. 10 shows the polarization curves of pure Ni coating and Ni–
Tafel slopes (βa and βc) were obtained from the intersection of cathodic TiO2 nanocomposite coatings in 1 M HNO3 solution. Passive layer
and anodic Tafel curve tangents using the Tafel extrapolation method. formation is observed in anodic region of polarization curves. This
Also, the corrosion rates (rcorr) and polarization resistance (Rp) are behavior is due to the formation of NiO passive film in acidic solution
calculated from these data. Table 1 represents the results obtained from [31]. Passive layer formation current density of Ni–TiO2 nanocompo-
polarization tests. It is obvious that with increasing of TiO2 nanoparticle site coatings (about 0.01 A/cm2) is less than that of pure Ni coating
content in coating the corrosion current decreases and the corrosion (about 0.02 A/cm2). On the other hand, passive layer formation
potential shifts to a more positive potential. The data clearly reveals the potential of Ni–TiO2 nanocomposite coatings (about 290 mv) is more
improvement of corrosion protection by TiO2 nanoparticles. Polariza- negative compare to that of pure Ni coating (about 480 mv). It can be
tion resistance (Rp) increases TiO2 nanoparticle content in coating. It concluded that the presence of TiO2 nanoparticles in Ni matrix coating
can be concluded that codeposited TiO2 nanoparticles in Ni matrix of aids in the formation of a passive layer. The passive current of Ni–TiO2
coating increases the corrosion resistance in salty, acidic and alkaline nanocomposite coatings (about 10− 3 A/cm2) is more than that of pure
solutions. Obviously, the TiO2 nanoparticles played a major role for Ni coating (about 10− 4 A/cm2). It means that passive layer corrosion
improving the corrosion protection in two mechanisms. Firstly, these rate of Ni–TiO2 nanocomposite coatings is more than pure Ni coating.
TiO2 nanoparticles act as inert physical barriers to the initiation and The presence of TiO2 nanoparticles in the coating surface disturbs the
development of defect corrosion, modifying the microstructure of the continuity of NiO passive layer and accelerates the destruction of
nickel layer and hence improving the corrosion resistance of the passive layer which causes the increasing of corrosion rate.
coating. Secondly, dispersion of TiO2 nanoparticles in the nickel layer
results in formation of many corrosion micro cells in which the TiO2 3.5. Electrochemical impedance spectroscopy (EIS)
nanoparticles act as cathode and nickel metal acts as anode because of
the standard potential of TiO2 more positive than nickel. Such corrosion 3.5.1. EIS tests in NaCl solution
micro cells facilitated the anode polarization. Therefore, in the presence The Nyquist diagrams in 0.5 M NaCl solution is shown in Fig. 11.
of TiO2, localized corrosion is inhibited, and mainly homogeneous The impedance spectra for Nyquist diagram was analyzed by fitting
corrosion occurs [29,30]. the experimental data to equivalent circuit model Fig. 12a. A good fit

Table 1
Corrosion characteristic obtained from potentiodynamic polarization measurement for pure Ni coating and Ni–TiO2 nanocomposite coatings.

Solution TiO2 (wt.%) Ecorr (mv) βc (mv/dec) βa (mv/dec) icorr (nA/cm2) Rp (MΩ cm2) rcorr (mpy)×10− 3

NaCl
0 − 595 79 70 9.42 1.713 0.110
3.9 −569 107 93 7.17 3.017 0.084
6.5 − 536 99 76 5.83 3.206 0.068
8.3 − 507 95 70 4.42 3.964 0.051

NaOH
0 −955 111 63 19.3 0.905 0.227
3.9 − 751 93 63 12.8 1.275 0.151
6.5 − 490 98 58 4.1 3.863 0.048
8.3 −442 103 70 2.1 5.628 0.024

HNO3
0 − 124 39 53 2.95 × 105 3.31 × 10− 5 3.47 × 103
3.9 −90 41 57 2.28 × 105 4.54 × 10− 5 2.68 × 103
6.5 −75 48 62 1.86 × 105 6.32 × 10− 5 2.18 × 103
8.3 −63 45 61 1.53 × 105 7.35 × 10− 5 1.79 × 103
P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810 3809

Fig. 13. Nyquist impedance diagrams for pure Ni coating and Ni–TiO2 nanocomposite
coatings in 1 M NaOH solution.

tances in NaOH solution are listed in Table 2. Rct increases with


increasing of TiO2 nanoparticle content in coating.

3.5.3. EIS tests in HNO3 solution


The Nyquist diagram in 1 M HNO3 solution is shown in Fig. 14. The
Nyquist diagrams show one depressed capacitive loop at the higher
frequencies followed by an inductive loop that is observed in the
lower frequencies. The equivalent circuit model is shown in Fig. 12c. A
good fit with this model was obtained with an average error of about
2.7%. L and Rl represent the inductance and inductance resistances.
Fig. 12. Equivalent circuits used for numerical fitting of impedance plots for, (a) NaCl The presence of lower frequencies Rl–L loop may be attributed to the
solution, (b) NaOH solution, (c) HNO3 solution. relaxation process obtained by adsorption species like H+ads. and
NO3−ads on the electrode surface [37,38]. The point of intersection
between the inductive loop and the real axis presents (Rs + Rct) [39].
with this model was obtained with an average error of about 1.8%. Rs Charge transfer resistances in HNO3 solution are listed in Table 2. It is
represents the solution resistance. The capacitive loop, Rct − CPE, can obvious that with increasing of TiO2 nanoparticle content in coating
be attributed to the charge transfer reaction. The constant phase the charge transfer resistance increases.
element, CPE, is introduced in the circuit instead of a pure double layer
capacitor to give a more accurate fit [32]. Rct represents the charge
transfer resistance whose value is a measure of electron transfer
across the surface and is inversely proportional to corrosion rate [33].
Charge transfer resistances in NaCl solution are listed in Table 2. Rct
increases with increasing of TiO2 nanoparticle content in coating.

3.5.2. EIS tests in NaOH solution


Fig. 13 represents the Nyquist diagrams in 1 M NaOH solution. The
Nyquist plots exhibit a capacitive loop at the higher frequencies
followed by a diffusion-controlled charge transfer at the lower
frequencies [34]. The diffusion process may indicate that the corrosion
mechanism is controlled not only by a charge transfer step but also by
the diffusion process. The equivalent circuit model for electrochemical
behavior is shown in Fig. 12b. The experimental data has a good fit
with this model with an average error of about 2.3%. Cdl and Zw
represent the double layer capacitor and Warburg impedance. The
corrosion rate is affected by the inter-diffusion of Ni2+ and OH− ions,
there should exist a Warburg diffusion element that reasonably Fig. 14. Nyquist impedance diagrams for pure Ni coating and Ni–TiO2 nanocomposite
describe the diffusion effect of ions [35,36]. Charge transfer resis- coatings in 1 M HNO3 solution.

Table 2
Charge transfer resistance for pure Ni coating and Ni–TiO2 nanocomposite coatings obtained from EIS data fitting by equivalent circuit model.

Solution Rct (kΩ cm2)

Coatings Pure Ni Ni–3.9 wt.% TiO2 Ni–6.5 wt.% TiO2 Ni–8.3 wt.% TiO2

NaCl 263 371 423 486


NaOH 203 383 491 634
HNO3 2 × 10− 3 3.3 × 10− 3 3.9 × 10− 3 4.3 × 10− 3
3810 P. Baghery et al. / Surface & Coatings Technology 204 (2010) 3804–3810

4. Conclusions [11] A.S. Hamdy, M.A. Shoeib, H. Hady, O.F. Salam, Surf. Coat. Technol. 202 (2007) 162.
[12] B.K. Prasad, J. Wear 252 (2002) 250.
[13] T. Watanabe, Nano-Plating Microstructure Control Theory of Plated Film and Data
Ni–TiO2 nanocomposite coating was deposited by conventional Base of Plated Film Microstructure, Elsevier, New York, 2004.
direct current electrodeposition process. The content of TiO2 nano- [14] D. Galvan, Y.T. Pei, J.Th.M. De Hosson, J. Surf. Coat. Technol. 201 (2006) 590.
[15] X.H. Chen, F.Q. Cheng, S.L. Li, J. Surf. Coat. Technol. 155 (2002) 274.
particles in coating is influenced by current density, stirring rate and [16] A. Ciubotariu, L. Benea, M. Varsanyi, V. Dragan, J. Electrochim. Acta 53 (2008)
TiO2 nanoparticle concentration in the plating bath. The maximum 4557.
TiO2 wt.% in coating is obtained at 30 g/l TiO2 nanoparticles in the [17] Z. Ahmad, Principles of Corrosion Engineering and Corrosion Control,, 1st ed.
Elsevier, London, 2006.
plating bath, current density at 5 A/dm2 and a stirring rate at 180 rpm.
[18] S.C. Wang, J. Wei, J. Mater. Chem. Phys. 78 (2003) 574.
The codeposited TiO2 nanoparticles were uniformly distributed into [19] N. Guglielmi, J. Electrochem. Soc. 119 (8) (1972) 1009.
Ni matrix improved corrosion resistance and mechanical properties of [20] C.S. Lin, C.Y. Lee, C.F. Chang, C.H. Chang, J. Surf. Coat. Technol. 200 (2006) 3690.
[21] A.M.J. Kariapper, J. Foster, J. Trans. IMF 52 (1974) 87.
coating. The hardness and wear performance of Ni–TiO2 nanocompo-
[22] J. Fransaer, J.P. Celis, J.R. Roos, J. Metal Finish. 91 (1993) 97.
site coating are improved by increasing of TiO2 wt.% in coating. With [23] L. Shi, C. Sun, P. Gao, F. Zhou, W. Liu, J. Appl. Surf. Sci. 252 (2006) 3591.
increasing of TiO2 nanoparticle content in the coating the corrosion [24] S.H. Yeh, C.C. Wan, J. Plat. Surf. Finish. 84 (1997) 54.
current decreases and the corrosion potential shifts to more positive [25] M. Surender, B. Basu, R. Balasubramaniam, J. Tribol. Int. 37 (2004) 743.
[26] F. Hou, W. Wang, H. Guo, J. Appl. Surf. Sci. 252 (2006) 3812.
values. The presence of TiO2 nanoparticles in Ni matrix coating aids in [27] G.E. Dieter, Mechanical Metallurgy, 1st ed. McGrow-Hill, London, 1988.
the formation of a passive layer in HNO3 solution but causes to [28] M. Neale, M. Gee, A Guide to Wear Problems and Testing for Industry, 1st ed.
accelerate passive layer degradation. The EIS spectra demonstrate that William Andrew, 2002.
[29] A. Abdel Aal, J. Mater. Sci. Eng. A 474 (2008) 181.
corrosion rate was controlled by charge transfer resistance. Rct [30] X.H. Chen, C.S. Chen, H.N. Xiao, F.Q. Cheng, G. Zhang, G.J. Yi, J. Surf. Coat. Technol.
increases with increasing of TiO2 nanoparticle content in coating 191 (2005) 351.
which is an indication of corrosion resistance increasing. [31] H.H. Strehblow, Passivity of Metals, Advances in Electrochemical Science and
Engineering, vol. 8, 2004, p. 274.
[32] J.R. Macdonald, W.B. Johanson, Theory in Impedance Spectroscopy, John Wiley &
References Sons, New York, 1987.
[33] A.M. Abdel-Gabar, B.A. Abd-El-Nabey, I.M. Sidahmed, A.M. El-Zayady, M. Saadawy,
[1] C. Subramanian, G. Cavallaro, G. Winkelman, J. Wear 241 (2000) 228. J. Corros. Sci. 48 (2006) 2765.
[2] D.S. Rickerby, A. Matthews, Advanced Surface Coatings: a Handbook of Surface [34] N. Perez, Electrochemistry and Corrosion Science, Kluwer Academic Publishers,
Engineering, Blackie, Glasgow, 1991. Boston, 2004.
[3] Y. Zhou, H. Zhang, B. Qian, J. Appl. Surf. Sci. 253 (2007) 8335. [35] L. Wang, J. Zhang, Y. Gao, Q. Xue, L. Hu, T. Xu, J. Scripta Mater. 55 (2006) 657.
[4] K.H. Hou, M.D. Ger, L.M. Wang, S.T. Ke, J. Wear 253 (2002) 994. [36] A. Bai, P.-Y. Chuan, C.-C. Hu, J. Mater. Chem. Phys. 82 (2003) 93.
[5] L. Benea, P.L. Bonora, A. Borello, S. Martelli, J. Wear 249 (2001) 995. [37] M.A. Amin, S.S. Abd El-Rehim, E.E.F. El-Sherbini, R.S. Bayyomi, Electrochim. Acta
[6] C.S. Ramesh, S.K. Seshadri, J. Wear 255 (2003) 893. 52 (2007) 3588.
[7] M.A. Baker, J. Surf. Coat. Technol. 201 (2007) 6105. [38] M.A. Veloz, I. González, J. Electrochim. Acta 48 (2002) 135.
[8] J. Li, Y. Sun, X. Sun, J. Qiao, J. Surf. Coat. Technol. 192 (2005) 331. [39] W.J. Lorenz, F. Mansfeld, J. Corros. Sci. 21 (1981) 647.
[9] Y.C. Wang, S.C. Tung, J. Wear 229 (1999) 1100.
[10] A. Abdel Aal, K.M. Ibrahim, Z. Abdel Hamid, J. Wear 260 (2006) 1070.

You might also like