Você está na página 1de 14

Applied Surface Science 401 (2017) 385398

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Role of surface roughness on corrosion and fretting corrosion


behaviour of commercially pure titanium in Ringers solution for
bio-implant application
Bose Sivakumar a,b , Lokesh Chandra Pathak a,b , Raghuvir Singh a,b,
a
Academy of Scientic and Innovative Research, CSIR Madras Complex, Chennai, 600113, India
b
CSIR-National Metallurgical Laboratory, Jamshedpur, 831007, India

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

Article history: Inuence of roughness (ra ) from 43 to 474 nm on corrosion and fretting corrosion of commercially
Received 2 September 2016 pure titanium (CpTi) was studied in the Ringers solution. The anodic polarization and electrochemi-
Received in revised form cal impedance spectroscopy (EIS) revealed the highest corrosion resistance of CpTi with ra 43 nm and
13 December 2016
correlated well with the surface energy (SE ). The highest potential drop associated with the fretting cor-
Accepted 4 January 2017
rosion is observed for CpTi with ra 43 nm followed by 474 nm; this is found to correspond with the worn
Available online 7 January 2017
out area. The fretting current density (ifretting ) is several order higher than obtained during the potentiody-
namic polarization (without fretting) study. Fretting corrosion manifested by the drop in electrochemical
Keywords:
Titanium
potential is simulated with high accuracy using fretting current density and an initial contact area. Fret-
EIS ting corrosion at an applied potential (+250 mV(SCE) ) is produced much larger fretting corrosion current
Polarization density than during the open circuit potential (OCP).
XPS 2017 Elsevier B.V. All rights reserved.
Fretting corrosion

1. Introduction texture on the implant suitable for the load bearing applications.
While signicant efforts have been directed to understand the tis-
Fretting corrosion plays a vital role in the biomedical assem- sue integration [10,11], studies to unveil the effect of roughness or
blies as it results in accumulation of the metal ions and wear texture on corrosion and more specically on the fretting corro-
debris near the tissues causing inammatory reactions which war- sion of titanium metal/alloys are rare. Most of the available studies
rant the re-surgery of patient [14]. Research carried out, so far, correlating the roughness with corrosion resistance are performed
is focussed to observe the effects of (i) microstructural variations on the ferrous alloys such as stainless steel in chloride containing
and fretting parameters such as fretting amplitude, load/contact solution. The 304 stainless steel with different surface roughnesses
pressure, frequency, and solution chemistry [15], and (ii) sur- induced by grinding with the varying grits of emery papers revealed
face engineering methods on the fretting corrosion behaviour the lowering of pitting resistance in chloride containing solution.
of titanium metal/alloys [6,7]. Roughness and texture (induced Erosion by the aqueous sand slurry is also found to change the
by the chemical or mechanical surface engineering techniques) pitting potential of stainless steel via surface roughness [1214].
are extremely important parameters for the bio-implants as they The deterioration in corrosion resistance of AISI 316 stainless steel
guide the growth of tissues on the surface. Rough surface con- due to the increased surface roughness is observed by Lage et al.
tains large area which is desirable for the tissue anchoring through [15]. They also demonstrated an improvement in the corrosion
an osseointegration process. Modication of the bio-implant is resistance by reducing the surface roughness. Another study car-
often considered based on the fact that the tissue in-growth is ried out on AISI 301 stainless steel showed that the growth of the
favoured on the surface with roughness up to 100 m [8,9]. Enor- metastable pits encounter more difculty on the smooth compared
mous research has been aimed to create the roughness and physical to the rough surface [16]. An increase in the roughness, however,
does not necessarily alleviate the pitting resistance; its role rather
depends on the alloys-environment types. For instance, roughness
Corresponding author at: CSE Division, CSIR-NML, Burmamines, Jamshedpur,
does not inuence the corrosion resistance of AISI 304 stainless
Jharkhand 831007, India.
steel in urban rain water [17]. An increase in the surface rough-
E-mail addresses: chemsiva1986@gmail.com (B. Sivakumar), ness of stainless steel 304 is shown to accelerate the passive layer
lokesh@nmlindia.org (L.C. Pathak), rsr@nmlindia.org (R. Singh). growth in urban rain water, but noticed to retard in the sodium

http://dx.doi.org/10.1016/j.apsusc.2017.01.033
0169-4332/ 2017 Elsevier B.V. All rights reserved.
386 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

chloride (NaCl) solution [17]. The corrosion rate of carbon steel polarization curves. The drop in potential is, however, caused by
in synthetic ground water, a non-passivating system, is reported disrupting the passive lm attributed to the fretting; hence, the
to increase with the surface roughness [18]. The study suggests use of ifretting would be more meaningful and practical. The current
that there exists an optimal value of the roughness, which may density during anodic polarization curves (without fretting) is actu-
be critical to minimize the corrosion rate of underground steel ally produced by the uniform surface (either active or passive) and
pipe [18]. A study performed in a simulated soil-extracted solu- consists of no galvanic couple unlike formed during the fretting cor-
tion indicates that the surface roughness enhances the corrosion of rosion. It, thus, cannot represent the current density that is obtained
pipeline steel [19]. While the literature unanimously agreed upon from the fretting induced galvanic couple. In the present investi-
that the roughness deteriorates the corrosion resistance of stain- gations, the measured fretting area (anode of the galvanic couple)
less steel in chloride containing solution, the same is not necessarily is though very small compared to the entire specimen area (cath-
true for the non-chloride containing solution or for several specic ode), yet it is much larger that the instantaneous fretting stroke
solution-material combinations. The only study, to the best of our (180 m). After analysing the fretting current and so the current
knowledge, reported the effect of roughness on titanium in 0.9% density, the potential drop has been simulated (using Tafel equa-
NaCl solution indicates no degradation in its pitting resistance [20]. tion). The worn out surface after fretting corrosion is considered as
The pitting resistance of Ni-Ti alloy, however, was shown to reduce an anodic area for the simulation; the anode is assumed to be par-
markedly by varying the roughness [20]. tially/poorly passivated or oxide lm is continuously disrupted (at
The literature emphasizing the effect of roughness on fretting the experimental fretting frequency) from its surface. This is unlike
corrosion is even scanty. Landolt et al. [21] observed that the earlier study which considered the innitesimal area covered by
change in roughness of the counter alumina sphere against passive the instantaneous fretting stroke. The attempts have been made
stainless steel resulted in the variation of anodic current during to unveil the effect of surface roughness on corrosion resistance
tribological experiments in 0.5 M sulphuric acid. At an applied using anodic polarization and EIS; fretting corrosion using the fret-
force of 5N, the increase in roughness from 0.02 to 4.25 m was ting set-up at an open circuit potential and applied potential in the
found to enhance the current density of stainless steel from 0.05 to Ringers solution (a simulated body uid).
0.18 mA/cm2 [21]. The hard oxide or debris produced by the wear
or corrosion may be risky and can further enhance wear rate of the 2. Experimental
base alloys. A study showed that the rough surface facilitates easy
escape of oxide particles/wear debris from the contact zone, dur- 2.1. Materials
ing the fretting corrosion, thus reduces the damage as compared
to the smooth surface [22,23]. Escaping of oxide or wear debris to CpTi, grade-2 was used as a substrate having chemical com-
the surrounding zone reduces the tangential force, thus decreases position (in wt.%): N-0.01; C-0.04; H-0.01; Fe-0.20; O-0.18; and
the fretting corrosion. Despite deliberate roughness (ra up to 2 m) Ti-balance (Akshat Steel, Mumbai, India). Specimens of 3 mm thick
created on the femoral head and femoral neck tapers of modular disc were cut from the 15 cm long CpTi rod of 2 cm diameter. The
hip joints made up of titanium to improve the tissue growth, no discs were abraded using the abrasive silicon carbide papers from
attempt has been made to study the inuence of roughness on the 100 to 2/0 grit sizes followed by the mirror nish with 0.5 m
fretting corrosion behaviour of titanium [24]. suspended alumina particles in aqueous solution. Different rough-
Present paper reports the results obtained from a study per- nesses were produced on the CpTi surface using the emery papers
formed on the corrosion and fretting corrosion of CpTi with ra of varied grit sizes such as 100, 400, and 2/0 grade and by polish-
varied from 43 to 474 nm in Ringers solution. Drop in potential ing with the Al2 O3 slurry. The specimens were then ultrasonicated
towards the cathodic direction during fretting corrosion is deter- using acetone for 10 min followed by cleaning with the distilled
mined experimentally; such potential drop is simulated further water.
using few empirical parameters. The potential drop during fretting
corrosion tests, though, has been reported by the earlier studies, yet 2.2. Roughness and contact angle measurement
it is far from the complete understanding. The potential being an
important electrochemical parameter and thermodynamic quan- The roughness of different surfaces was measured using 2D
tity indicates the fretting corrosion damage, which needs further (dimensional) surface prolometer (Taylor Hobson 2 make). A 3D
elaboration to advance the understanding [25,26]. Celis et al. [27] surface prolometer (DektakXT model) Bruker make was employed
described the potential drop via change in the standard electrode to determine the fretting wear depth after fretting corrosion exper-
potential, Eo (of Nernst equation) due to the variation of pressure at iments. The contact angle measurements based on the sessile drop
the fretting contact; which otherwise is reported at 1 atmospheric method using SCA20 software (makeData Physics, model15EC)
pressure and a unit ion activity. The material deformation caused were performed to calculate the surface energy of CpTi with the var-
by the fretting load is expected to alter the enthalpy and thereby ied ra ; droplet of the double distilled water was utilized for 5 s at
the Gibbs free energy including Eo . These are the possible factors, room temperature (20 2 C). Minimum seven drops were used for
in addition to hydrodynamics and material-solution combination each measurement and their average is reported in the manuscript.
(passive or active type), which lead to a drop in potential during
the fretting corrosion. While these could be the underlying rea- 2.3. Anodic polarization and electrochemical impedance
sons, quantitative description and possible simulation may be a spectroscopy
better assessment of drop in potential. The quantitative analysis
of the potential drop during fretting corrosion at OCP condition To investigate the corrosion behaviour, anodic polarization
has been recently proposed by Vieira et al. [26]. They considered and AC impedance methods were performed using potentio-
galvanic couple where the area under the fretting stroke acts as stat/galvanostat frequency response analyser, Gill AC Bi-STAT
anode and surrounding area as cathode and used Tafel equation model, ACM instruments, UK. A three electrode system contain-
to calculate the potential drop. This has been attempted for the ing CpTi discs as working electrode; the graphite, and saturated
tribo-corrosion of aluminium alloy in 0.05 M NaCl and 0.1 M sodium calomel electrode (SCE) as counter and reference electrodes respec-
nitrate (NaNO3 ) solutions. The anodic current density used to simu- tively were used. These tests were conducted in the Ringers
late the potential drop is chosen arbitrarily and the best tting value solution at room temperature (20 2 C). The Ringers solu-
was compared with the current density obtained from the anodic tion was prepared by adding (in g/L) 9Soidum chloride (NaCl),
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 387

Fig. 1. Schematic of the fretting corrosion set-up.

0.24Calcium chloride (CaCl2 ), 0.43Pottassium chloride (KCl), and mer material) and designed in such a way that the only 2 cm2 of
0.2Sodium bicarbonate (NaHCO3 ) in distilled water, maintained at working electrode is exposed to the experimental solution. The
pH7.3. Anodic polarization curves were obtained by scanning the fretting counterparts (alumina ball and CpTi) were immersed in
specimen from 250 mV(SCE) (with respect to OCP) to the anodic the solution. The rear of the working electrode was connected to
region up to +3000 mV(SCE) at a scan rate of 100 mV/min. The poten- the potentiostat/galvanostat (ACM instruments, UK make); SCE and
tial sweeping up to such a high potential value is considered as graphite electrodes were employed as reference and counter elec-
CpTi normally exhibits large (+ve) breakdown potential. All the trodes respectively (Fig. 1).
tests were repeated three times to ensure the reproducibility of The OCP (without fretting) measurement was initiated 1 h
the data. EIS test at OCP was carried out within the frequency (hour) before the fretting was started in order to attain a steady
range 100000.01 Hz at 10 mV amplitude. The ZSimpWin software state potential by the CpTi in Ringers solution. After the fretting
was used for analysing the EIS results according to the suggested was put ON, the OCP and current measurements were continued.
electrochemical equivalent circuit (EEC) t model. While the OCP and current were continuously monitored, the fret-
ting was allowed to run for 15 min in 5 cycles (15 min ON and
2.4. Fretting corrosion behaviour 15 min OFF) for different durations. Similar experiments were car-
ried out with the fretting cycles of varied durations such as 30
The fretting corrosion behaviour of CpTi with different sur- and 60 min, in separate experiments. The fretting parameters used
face roughnesses (from 43 to 474 nm) was monitored at OCP were: 3N load, an oscillating frequency of 5 Hz, and linear displace-
and applied anodic potential (+250 mV(SCE) ) in Ringers solution at ment of 180 m. All the fretting corrosion tests were repeated for
20 2 C, using fretting corrosion set-up [5,6]. The test assembly three times using a fresh specimen to ensure the reproducibility.
was fabricated by the Wear and Friction Tech, Chennai, India, as The worn surface morphology was studied by the optical (Leica,
schematically demonstrated in Fig. 1. The fretting corrosion set-up Germany make) and scanning electron microscopy (SEM, Nova
with ball-on-at contact method, where the alumina ball of 8 mm Nano SEM 430, Netherland make). The fretted area was determined
diameter with ra 20 nm (obtained from Salem Speciality Ball Com- by using optical microscope equipped with the ImageJ software.
pany Inc, USA) was used as a counter body that fretted against the
stationary CpTi specimen. A metal on ceramic contact is frequently 2.5. X-ray photoelectron spectroscopy (XPS)
considered in the load bearing implant assembly. The applied nor-
mal force using spring assembly was controlled by the load cell, The analysis of oxide formed on the CpTi surface (of differ-
while the fretting displacement was monitored by the variable fre- ent roughnesses) after 24 h of exposure to the Ringers solution
quency drive motor (linear actuator). The load to the alumina ball, was studied using XPS (SPECS, Germany make) with Alk
held by the 80 mm long Teon holding device, was transferred by (hv1486.6 eV) and Mgk (hv1253.6 eV) electron sources. Vac-
the spring assembly. The entire set up was placed on a steel plate uum of the chamber was maintained at 109 Torr during the
for good stability and operated by the control panel. The fretting acquisition of XPS spectra. The deconvolution of XPS peaks was
corrosion cell was fabricated with the Delrin (Acetal homopoly- performed using CASA XPS software. The C1s binding energy (BE)
388 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

3000 (a) Exp- ra 474 nm


250
Fit- r a 474 nm
2500 Exp- ra 116 nm
ra 474 nm 200 Fit - ra 116 nm
Potential (mV/SCE)

Exp- ra 177 nm

- Z'' (kohm cm )
2000 ra 177 nm

2
Fit- ra 177 nm
ra 116 nm 150
1500 60

ra 43 nm

- Z'' (kohm cm2)


100
1000 30

500 50
Exp- ra 43 nm
0
Fit - r a 43 nm 0 30 60
0 0 Z' (kohm cm2)

0 50 100 150 200 250


-500 2
Z' (kohm cm )
-8 -7 -6 -5 -4 -3
10 10 10 10 10 10 (b) 104
Exp- ra 177 nm Exp- ra 43 nm
2 Fit- r a 177 nm 90
Current density (A/cm ) Fit - ra 43 nm
3
10 75
Fig. 2. Polarization behaviour of CpTi with the surface roughness ranging from ra

- Phase angle (degree)


2

Z (kohm cm2)
43 to 474 nm in Ringers solution. 10 60

1 45
peak position at 284.6 eV generated due to the hydrocarbon layer 10
was used as a reference for XPS analysis. The change in the atomic 30
0
10 Exp- ra 474 nm
concentration of oxygen (O1s) component was calculated using Fit- ra 474 nm 15
the sensitivity factor of 2.85. The mixture of the Gaussian and -1 Exp- ra 116 nm
10
Lorentzian forms was used for the curve tting and the Shirley Fit - r a 116 nm 0
method was used for the background corrections. -2 -1 0 1 2 3 4
10 10 10 10 10 10 10
Frequency (Hz)
3. Results and discussion
Fig. 3. (a and b) Nyquist and Bode impedance with phase angle plots of CpTi with
3.1. Anodic polarization and EIS behaviour the surface roughness ranging from ra 43 to 474 nm in Ringers solution.

The anodic polarization curves of CpTi with different ra


(43474 nm), obtained in the Ringers solution, are shown in Fig. 2.
These indicate typical passive-transpassive behaviour. The elec-
trochemical parameters such as Ecorr (corrosion potential), icorr
(corrosion current density) and corrosion rate were obtained from
the Tafel extrapolation method, while ipass (passive current density)
was determined from the potentiodynamic polarization curves.
Various parameters calculated from the anodic polarization curves
are listed in Table 1. The Ecorr values were observed to be 383,
257, 344, and 333 mV(SCE) for the surface with ra 43, 116,
177, and 474 nm, correspondingly. The CpTi demonstrated two Fig. 4. EIS tting model used for CpTi with different surface roughness.
passive regions ranging from the potential 100 to 1300 mV(SCE)
and >1750 mV(SCE) with corresponding passive current densities
designated as ipass1 and ipass2 respectively (Table 1). This is due CpTi surface with different roughnesses (ra 43474 nm) show the
to the chemical change occurring in the passive lm with the two depressed semicircles corresponding to two time constants.
increasing electrochemical potential. While the passivation at a The at region in the Bode impedance curves <100 Hz corresponds
lower potential range is accounted for the predominance of TiO2 , to the electrolyte resistance (<0.1 k cm2 ), while a linear increase
the latter could be due to the sub-oxide/hydroxide formation in the impedance between 100 and 10000 Hz indicates increase in
such as TiO(OH)2 [28]. Similar transition in the passive region at the capacitive behaviour of passive lm. The EEC tting considered
1300 mV(SCE) has been reported earlier and change in the nature duplex passive lm consists of compact (inner) and porous (outer)
of oxide lm (from TiO2 to TiO and Ti2 O3 ) is held responsible for the layers, which is consistent with the earlier reports [30,31]; various
same [29]. The surface with the ra value of 43 nm is corroded with parameters of these layers are presented in Table 2. The good-
the lowest icorr (41.1 04 nA/cm2 ), compared to the higher ra sur- ness of tting, indicated by chi-square values (2 ), of impedance
faces. The ipass1 of ra 43 nm surface is one third (3.8 0.3 A/cm2 ) of curves is obtained in the range from 103 to 104 . A duplex layer
the CpTi with roughness ra > 43 nm. The icorr and corrosion rate are model represented as Rs (Rp (Qp (Rc (Qc ))) is shown in Fig. 4, where
observed to decline with an increase in the ra from 116 to 474 nm. Rs is solution resistance; Rp and Qp are the resistance and constant
Likewise, the ipass1 in the passive region <1.3 V(SCE) also followed phase element of an outer porous layer; Rc and Qc are the resis-
the similar trend. tance (charge transfer resistance) and constant phase element of
To support the corrosion behaviour as obtained from the anodic an inner compact layer. The Q is used instead of pure capacitance
polarization, electrochemical impedance study of CpTi surfaces (C) due to the deviation from an ideal capacitor where impedance

with different ra values was carried out. The EIS plots, in both the of Q (ZQ ) is equal to (C(j)n )1 (in 1 cm2 sn ); j is ( 1), = 2f
formats e.g. Nyquist (Z vs Z) and Bode impedance (|Z| vs fre- is angular frequency (in radian/sec), f is frequency (in Hz) and n
quency) with phase angle are illustrated in Fig. 3(a and b). The is an exponent ranges from 1, 0 to 1; 1 being pure capaci-
gure includes the data obtained using the EEC model (as shown tor, 0 a pure resistance, and 1 represents a pure inductance.
in Fig. 4) tting. The Nyquist curves in Fig. 3(a) (magnied view) of Among various surface roughnesses, ra 43 nm shows the highest
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 389

Table 1
Surface energy and electrochemical parameters obtained from polarization curves of CpTi with different surface roughnesses.

Surface Surface energy a (mV/decade) c (mV/decade) Average icorr Average ipass1 Average ipass2 Average corrosion
roughness (nm) (mJ/m2 ) (nA/cm2 ) (A/cm2 ) (A/cm2 ) rate 103 (nm/h)

474 48 22.42 1.7 200 140 52.0 03 9.61 0.12 63 04 57.42 3.4
177 7 30.35 2.2 208 136 70.2 10 12.4 0.44 70.1 03 74.20 10.9
116 9 38.43 1.8 215 116 105.0 13 13.2 0.30 496.4 33 107.0 13.3
43 5 15.94 2.7 223 148 41.1 04 3.8 0.30 82.2 02 40.63 4.4

Table 2
Equivalent circuit tting parameters obtained from the EIS results of CpTi with different surface roughnesses (43474 nm).

Surface Rs ( cm2 ) Qp x 105 (1 cm2 sn ) n1 Rp 103 ( cm2 ) Qc x 106 (1 cm2 sn ) n2 Rc x 105 ( cm2 ) Chi-square
roughness (nm)

474 48 54 3.34 0.94 3.79 4.60 0.78 4.24 1.70 103


177 7 52 4.22 0.93 14.27 2.15 0.95 2.70 3.78 103
116 9 51 4.35 0.90 81.28 0.60 0.90 1.71 9.65 104
43 5 50 1.66 0.94 5.49 2.41 0.97 10.73 4.46 103

resistance of the inner layer, Rc 10.73 105  cm2 due to the for- growth, an occluded area is pre-requisite. The increasing rough-
mation of less defective and impermeable TiO2 layer. The larger ness offers more restricted geometry, due to the occurrence of a
diameter of semicircle associated with the ra 43 nm, on Nyquist large number of grooves/occluded sites favourable for the pitting
plot, supports the higher resistance of inner layer, compared to the or crevice corrosion. The roughness may inuence the anodic polar-
CpTi of ra > 43 nm. The Rc is observed to increase from 1.71 105 ization curve either during the active corrosion region (e.g. carbon
to 4.23 105  cm2 with an increase in roughness from 116 to steel in chloride solution) or during the passive state (e.g. stainless
474 nm. The resistance of the outer layer, though, decreased from steel in chloride solution). The variation in roughness may result in
81 103 to 3.8 103  cm2 with ra from 116 to 474 nm, the resis- increased corrosion rate (when alloy is under an active corrosion
tance of CpTi is predominantly results from the inner compact layer. process) or early breakdown of the lm/reduction in the pitting
The outermost lm in the titanium is shown to be defective and vul- potential (when alloy exhibits passive zone). The roughness, for
nerable to dissolution [32,33]. The Bode phase angle of CpTi with instance, is clearly shown to reduce the pitting potential of stainless
different ra exhibits a phase angle maxima of 80 on a wide fre- steel in chloride containing solution [12].
quency range (200.1 Hz), as depicted in Fig. 3(b). The higher phase The CpTi, however, is well known to afford the high pitting and
angle of the surface with ra < 116 nm is an indicative of the higher crevice corrosion resistance (up to >2.5 V(SCE) in chloride contain-
passive lm resistance. ing solution) due to the formation of an extremely impermeable
Both the methods (anodic polarization and EIS) inferred rela- TiO2 lm. The electrolyte chemistry suitable for the initiation of
tively a better corrosion resistance of CpTi with ra 43 nm. However, pitting/crevice corrosion (in CpTi) would, therefore, be achieved at
the icorr and ipass1 (at <1.3 V(SCE) ) were noticed to decrease from 105 a chloride concentration much higher than is needed for the stain-
to 52 nA/cm2 and 13.29.61 A/cm2 with the surface roughness less steel. The roughness effects on lowering the pitting potential of
increased from 116 to 474 nm respectively. These values may fur- CpTi surface may thus not be expected in the Ringers solution. Con-
ther be lowered considering the true surface area which increases trarily, a slight increase in the corrosion resistance of CpTi beyond
(more than the apparent area) with the increase in ra . The results 116 nm surface roughness, obtained in this study, indicates it to
infer that there may be a critical roughness value (such as ra 43 nm) be useful. This fact is supported by the earlier studies that evi-
below which the corrosion resistance of CpTi improves, as also denced no degradation in corrosion resistance of Ti alloy with
reported by the earlier researchers [34]. The possible reasons for the increase in the roughness [34]. They rather found a decline
such an observation may be: (i) the surface with ra 43 nm may in corrosion resistance of Ni-Ti alloy due to increase in the sur-
not form a convection restricted geometrical surface (with mini- face roughness. Observations in the literature and present study
mum undulations and more open types) which is pre-requisite to thus indicate that the different alloy-solution may behave differ-
aggravate the solution chemistry for pit formation and (ii) the sur- ently with respect to corrosion resistance under the inuence of
face (with ra 43 nm) is nearly an ideal to provide a template for roughness. The improvement in the corrosion resistance of CpTi
the formation of least defective passive lm so that the pit initiation (ra > 116 nm) may be attribute to thickening of the TiO2 layer driven
becomes difcult. It is further envisaged by the SE calculations using by the increase in surface roughness, which may be explained by
contact angle measurement (shown in Table 1), that the surface the fact that the rougher surfaces become electrochemically more
with ra 43 nm contained the lowest energy (15.94 mJ/m2 ) and active. It is evident that the difference in Volta potential using the
thus the most stable or least reactive among all the surfaces. Rela- Kelvin probe force microscopy, a physical quantity reciprocal to
tively a higher contact angle measured on this surface led it to be a the electron work function, increases with the surface roughness
more hydrophobic (repulsive to the water molecule adsorption and of carbon steel [18]. It has been shown earlier that the peaks on
humidity) and thereby possesses a higher corrosion resistance than the rough surface have lower electron work function compared to
other surfaces. This (contact angle) is followed by the CpTi surface the valleys; it means less energy is required to eject an electron
with ra 474, 177, and 116 nm which corresponds well with their and therefore peaks are more electrochemically active [18]. The
corrosion resistance. The SE values were calculated using SE = SEvl difference between the electronic behaviour of the two locations
cos , described elsewhere [35], where SEvl, the surface energy (peak and valley) hence would lead to an increased driving force
between water and air at an ambient condition is 72.8 mJ/m2 [35] for corrosion and hence thickening of the passive lm (as on the
and  is the static contact angle measured experimentally. surface of ra 474 nm). This is also supported by the measured SE
Generally, the growth of the pit/crevice in chloride containing which decreased from 38 to 22 mJ/m2 with the roughness varied
solution is governed by the diffusion of metal ions from within the from 116 to 474 nm. The lowest SE of CpTi with ra 43 nm followed
pit to the external solution and therefore for a sustained pit/crevice by 474 nm depicts the passive lm stability either due to being a
390 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

O1s, ra- 474 nm O1s, ra- 43 nm


O2-
O2-
OH- and defect

Intensity (a.u)
Intensity (a.u)

OH- and defect

540 535 530 525 540 535 530 525


B.E. (eV) B.E. (eV)

Ti2p 2p3/2
Intensity (a.u)

2p1/2

470 465 460 455


B.E. (eV)

Fig. 5. The XPS spectra of O1s components and Ti2p spectra of CpTi with ra 43 and 474 nm.

defect free (ra 43 nm) or the higher thickness (ra 474 nm). The pas- ing part of the BE peak position of O1s shows shoulder component
sive region >1.75 V(SCE) though indicates a higher current density at 532.50 eV which corresponds to the OH addition to Ti (as Ti-
(62 A/cm2 ) of ra 474 nm than that at <1.3 V(SCE) (<10 A/cm2 ), OH) and non-stoichiometry of TiO2-n . The concentration (at.%) of
it remains constant with further increase in the potential. Such OH on the surface of different ra values 43, 116, 177, and 474 nm
transition appears to be inuenced by the roughness; the higher were obtained to be 39.78, 44.94, 49.25, and 46.11 respectively.
roughness (474 nm) caused a loop with the higher peak current The corresponding O2 concentration was 60.22, 55.05, 50.77, and
density (1 mA/cm2 ) as compared to other. This is possibly due 53.88. It supports the hypothesis of the convective restricted geom-
to the fact that the slightly more aggressive solution chemistry in etry, discussed as above, that oxygen depleted with the increase
the groove needs more current to transform the oxide product to in roughness. The ratio of OH /O2 is (0.66, 0.82, 0.97, and 0.86)
be more protective. It is similar to that the higher primary passive found to increase with the roughness which indicates the increase
current densities (in active corrosion region of polarization dia- in hydration of the lm. The increase in hydroxyl ion concentration
gram) are required to passivate the stainless steel with the lower on the surface is often considered to be useful for the bio-implant
chromium contents [36,37]. application as the same can accelerate the hydroxyapatite forma-
tion on the surface.
3.2. XPS analysis of passive lm formed after OCP exposure

The analysis of the passive lm formed on CpTi of ra values


43474 nm, after 24 h of exposure in Ringers solution (at OCP) 3.3. Fretting corrosion at open circuit potential
was carried out using XPS; the spectra obtained are shown in
Fig. 5. The Ti2p XPS spectra show two peaks such as Ti2p3/2 The fretting corrosion of CpTi with surface roughness varied
and Ti2p1/2 due to the spin-orbit coupling. The peak positions of from ra 43 to 474 nm was studied as a function of time in Ringers
Ti2p1/2 and Ti2p3/2 from the CpTi of all the roughnesses appeared solution as described earlier. The variation in OCP/drop in poten-
at BE 464.4 eV and 458.7 eV respectively. These peaks suggest tial and current density with the time during fretting ON and OFF
Ti exists in Ti4+ oxidation state as TiO2 . The peak positions of BE are illustrated in Fig. 6(ac). The fretting ON and OFF were per-
associated with the Ti2p1/2 and Ti2p3/2 obtained are similar to formed to mimic the micro-movements such as walking and rest
as reported elsewhere [38,39]. The peak energy separation (BE) states of the patients and its consequences on the contact interface
obtained between Ti2p1/2 and Ti2p3/2 is 5.7 eV which is shown to of the hip and knee joints. The resultant changes in OCP/drop in
vary from 5.6 to 6 eV [38,39]. In an attempt to investigate the O1s potential and current obtained manifest the surface damage due to
peak from the TiO2 layer, two components, main and shoulder are the fretting corrosion. Monitoring of the two parameters (potential
observed on all the CpTi surfaces appeared as a main component at and current) describes the changes in damage pattern at the con-
530 eV, indicating lattice oxygen atom existing as O2 oxidation tact zone brought out by the thermodynamic and kinetic factors
state. This further conrms the presence of TiO2 layer. The remain- associated with the electrochemical reaction.
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 391

-100 (a) Oxidaon of


OFF OFF OFF OFF
oxide covered
-200
EoO2/H2O CpTi surface
Potential (mV/SCE)

ON ON ON ON ON
-300 ioO2/H2O Ecorr of oxide Oxidaon of
covered CpTi

Potential, V
before bare CpTi
-400 fretting
Ecorr of bare CpTi

fretting
ends
-500 EoTi/Ti4+ icorr due to oxide icorr due to
covered CpTi bare CpTi
-600
ra 474 nm ra 116 nm
ioTi/Ti 4+
-700 (a) ra 177 nm ra 43 nm
0 2500 5000 7500 10000 12500
Time (s) Current density (arbitrary unit)

-200 OFF OFF OFF


OFF (b) (O2, H+ reducon on CpTi
ON ON surface surrounding freng
-300 ON ON
ON Estac zone zone)
Potential (mV/SCE)

-400 fretting ends


before fretting

Potential, V
-500 Potenal during
freng
-600

-700 Oxidaon on the surface


r a 47 4 nm r a 116 nm Efreng zone
under freng
-800 (b) r a 17 7 nm r a 43 nm
0 5000 10000 15000 20000 25000
Time (s)
Current density (arbitrary unit)
2
Current density (nA/cm2)

Fig. 7. (a and b) Schematic showing the variation in (a) potential and current due to
r a 474 nm , 15 m in frett ing 'ON ' the oxide coverage on the CpTi surface and (b) potential and current with increase
1 in fretting area/time; dashed lines represent the changes occurred as cathode area
decreases with the progress in fretting.

0
equation ETi/Ti++ = 1.63 0.059 log Ti++ ) considering the lowest
Fretting ON Fretting OFF
corrosion state of CpTi when Ti++ concentration is 106 ppm
-1
[40]. The corresponding reduction reactions governing the corro-
sion of titanium (in studied solution) are oxygen and hydrogen
(c) evolutions. The redox potential for oxygen and hydrogen reduc-
-2
0 200 400 600 80 0 1000 tion reactions on CpTi surface in Ringers solution at pH 7.3 is
Tim e (s) estimated to be +0.557 V(SCE) and 0.673 V(SCE) respectively (using
EO2/H2O = 1.23 0.059 pH and EH+/H2 = 0.059 pH) [40,41].
Fig. 6. (ac) Fretting corrosion behaviour of CpTi with ra 43474 nm at OCP in
Ringers solution: (a) 15 min fretting, (b) 60 min fretting and (c) representative The value of corrosion potentials obtained (257 to
galvanic current from CpTi with ra 474 nm during fretting ON for 15 min. 383 mV(SCE) for CpTi surface of different ra values) lie between
+0.557 V(SCE) (EO2/H2O ) and 1.518 V(SCE) (ETi/Ti++ ) according to the
mixed potential theory or as a result of the combination of all the
3.3.1. Thermodynamic description of potential drop redox reactions in the system. The corrosion potential of titanium
The fretting ON during experiment is reected by the sud- is established by the oxygen reduction reaction, as being the more
den drop (within 10 s) in electrochemical potential towards more positive (+ve) than hydrogen reduction. The corrosion potential
cathodic (negative) side accompanied by the concurrent rise in will be inuenced further by the kinetic parameters associated
anodic current. The corrosion potential established on the bare CpTi with the chemical species in the solution that forms oxide. The
and oxide covered CpTi is shown by the schematic in Fig. 7(a); fretting at any instance shall cause the removal of passive lm
this represents the increase in OCP (towards more positive val- accompanied by the drop in potential below the corrosion poten-
ues) and decrease in the corrosion current as a result of the higher tial up to 1.518 V(SCE) . The drop in OCP during fretting corrosion,
Tafel slopes on the oxide covered CpTi. The negative drop in the in the present experiments, is found to be as low as 0.759 V(SCE) .
OCP during fretting corrosion is manifested by the removal of Such potential values are never reported for CpTi during the
existing passive oxide lm formed during the fretting OFF and static OCP measurements (without fretting), even on the freshly
exposure of the fresh CpTi surface to the Ringers solution. The prepared surface in Ringers solution. This is due to the fact that the
maximum drop in potential due to the fretting corrosion is, how- CpTi has high afnity to oxygen; surface even in freshly prepared
ever, guided by the oxidation of titanium and reduction of oxygen condition gets quickly covered with the nanometres thick oxide
and hydrogen in the Ringers solution. The redox potential of CpTi layer. This indicates that the CpTi may realize such a large negative
(ETi/Ti++ ) in the present experimental solution is 1.518 V(SCE) (using potential when its surface is free from any oxide. As soon as the
392 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

0
Potential (mV/SCE)

-100
ra 474 nm

-200 ra 43 nm

-300

-400

-500
0 10 20 30 40 50
Fretting area (cm2 x 10-4)

Fig. 8. Illustrates the drop in potential (cathodic) with the increase in fretting area
of CpTi with ra 43 and 474 nm.

fretting disrupts the passive lm, the corroding system becomes


galvanic with two unequal zones (small anode under the fretting
and cathode outside the fretting zone) establishing two different
Fig. 9. Worn out surface area of CpTi with the surface roughness ranging from ra 43
electrochemical potentials; this difference in potential accelerates
to 474 nm and counter surface (alumina ball) after fretting corrosion for 300 min in
the corrosion. This is shown by the schematic in Fig. 7(b); where Ringers solution.
the overall cathodic reactions are represented on the passive
CpTi (outside the fretting zone) and anodic oxidation of CpTi on
the fretting zone. As soon as the galvanic couple forms, a large ball is close to 1 and do not vary signicantly with the roughness or
peak on the current versus time curve is appeared just after the fretting time and hence ensures the validity of the results. Owing
fretting ON, as shown in Fig. 6(c). The electrochemical potential to the higher initial contact area between the alumina ball and CpTi
monitored during the fretting corrosion is thus established by the with ra 43 nm, smaller time is needed to level off (by fretting action)
two zones forming a galvanic couple. This new OCP after fretting followed by the damage below the surface (in subsurface region).
corrosion must be closer to the redox potential of the bare titanium This results in the adhesive galling and induces a greater fretting
(Eo Ti/Ti4+ ). As fretting progresses, the anodic zone broadens which depth, as shown in Fig. 10(a) (3D surface topography). The material
inuences and re-establishes the electrochemical potential. The transfer from titanium to the alumina ball at fretting zone is sup-
potential further decreases to more ve values with increase in ported the adhesive galling wear, as shown in Fig. 10(c). However,
the fretting area, as demonstrated by Fig. 8. This has been shown fretting on the surface of the highest roughness (ra 474 nm) is facil-
for the most ideal surface, with ra 43 nm and the roughest surface itated by the abrasive wear due to the presence of large peaks and
with ra 474 nm. The decrease in absolute potential drop with valleys and so yielded a larger worn area. The depth of the fretting
the increase in the fretting area/time (as shown in Fig. 8) is in was lesser on the surface of ra 474 nm, compared to 43 nm surface,
compliance with the schematic illustrated in Fig. 7(b). However, as shown by 3D surface topography in Fig. 10(b). The higher fretting
for longer fretting period, the oxide may get cladded to the surface area on the surface of ra 43 and 474 nm than that on the 116 and
and the effective anode area is reduced, this may then shorten the 174 nm corresponds with the larger potential drop as mentioned
magnitude of potential drop. The highest drop in potential due to above.
fretting corrosion has been observed for the CpTi with ra 43 nm
followed by 474 nm. The drop in potential on the former could be 3.3.2. Kinetics of fretting corrosion
related to the more surface area under the contact during fretting The galvanic current during the fretting ON corresponding to
leading to depassivation and the fresh exposure of bare CpTi to the cathodic shift of OCP on CpTi with various ra values, is shown
the solution. This was followed by the surface with ra 474 nm. in Fig. 11(ad). The fretting ON is manifested by the rise in current
Generally, the least drop in OCP was noticed for the CpTi surface which oscillates with larger amplitude compared to that during
with ra 116 nm (155 mV(SCE) ). While the fretting brings drop in the fretting OFF or static OCP measurement. The average galvanic
potential about certain mean value, the potential transients about current is found to increase after fretting ON cycle from about
the mean (during fretting) are indicative of depassivation of the 0.285 to 0.325 nA/cm2 . The current oscillates, corresponding to
lm during fretting oscillation. the frequency of the fretting movement, about the mean value;
the lower limit of the current drop appears similar to the current
3.3.1.1. Fretting area and roughness. The worn out surface area of during the static OCP or fretting OFF (Fig. 11(b and c)). This indi-
CpTi and alumina ball (counter surface) after various fretting dura- cates that the reformed passive lm following the fretting motion
tions is found to be roughness dependent, as illustrated in Fig. 9. is equally resistant to the lm that formed during the static OCP
The fretted area is observed to increase with the fretting time. The (fretting OFF cycle). However, in few cases, the lower value of cur-
largest worn out area was observed for the titanium surface with rent drop remains slightly higher than the current during static OCP
ra 43 nm; it is 35.8 104 and 45.8 104 cm2 after subjected to (Fig. 11(a)) suggests that the reformed passive lm does not pro-
the fretting corrosion for 75 and 300 min respectively. The smallest vide the same corrosion resistance as offered by the lm formed
worn out area was observed for the CpTi with the ra value of 116 nm, during static OCP.
for instance it is 13.7 104 and 25.4 104 cm2 after 75 and The change in current values obtained (during fretting ON or
300 min of the fretting corrosion. The worn out area of the counter OFF) are drawn from the corrosion of entire surface (area under
body (alumina ball) follows a trend similar to as determined for the fretting zone + surrounding the fretting zone). The current rise after
CpTi surface. The ratio of the worn out areas of titanium to alumina fretting ON is, however, accounted for the depassivation of small
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 393

Fig. 10. (ac) 3D surface topography depicting the fretting depth of CpTi (at OCP) with (a) ra 43 nm, (b) ra 474 nm and (c) fretting wear morphology on counter surface
(alumina ball) by optical microscope; adhesion of the wear debris is indicated by the arrows.

area under the contact interface. Thus the total current, Itotal , after where a is the radius of the spherical contact, P is normal load 3N,
fretting ON is made up of two components: Br is radius of the ball 4 mm, and Yreduced is the reduced modulus.
The reduced modulus was obtained using the following equation:
Itotal = IOCP + Ifretting (1)
1
Yreduced =     (4)
Ifretting = Itotal IOCP (2) 1(alumina )2 1(Ti )2
Yalumina
+ YTi

where IOCP is the current obtained during static OCP measurement


where alumina and Ti are the poisons ratio of alumina ball and
and contributed by the area outside the fretting region (under pas-
CpTi respectively, Yalumina and YTi indicate the Youngs modulus of
sivation); Ifretting is the absolute current increased due to the fretting
alumina ball and CpTi respectively. The values of various constants
accounted for a very small contact area. Such a small contact area
are taken from the literature as: alumina 0.238; Ti 0.37; Yalumina
can be determined by using the following equation described for
313 GPa; YTi 105 GPa [43].
Hertzian contact theory [42] for the two elastic bodies (ball and at
The contact radius determined between the alumina ball and
surface) as:
CpTi using the above relation is 6.5 105 cm2 and considered to

3
be the fretting area at the start of fretting (time t0 ). This contact area
a= (3PBr ) /4Yreduced (3) is responsible for the large initial jump in the fretting current den-
394 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

0.32 (a) ra 474 nm (b) ra 177 nm

Current density (nA/cm2)


0.32

Current density (nA/cm2)


0.31
0.31

0.30
0.30

0.29
0.29
Fretting ON Fretting OFF
0.28 Fretting ON Fretting OFF

4000 5000 6000 7000 4000 5000 6000 7000


Time (s) Time (s)

(c) ra 116 nm (d) ra 43 nm


0.32
Current density (nA/cm2)

Current density (nA/cm2)


0.31

0.31
0.30

0.30
0.29
Fretting ON Fretting OFF
0.29 Fretting ON
Fretting OFF
4000 5000 6000 7000 4000 5000 6000 7000
Time (s) Time (s)

Fig. 11. (ad) Fluctuations in the galvanic current after fretting ON from CpTi with the various roughness values (ra 43474 nm).

ting corrosion. The drop in OCP which is equivalent to the cathodic


overpotential due to polarization resulted from the fretting can be
simulated by the Tafel equation as:

 = ac + bc log ic (5)

where  is overpotential, ac = bc log io ; bc is Tafel constant; io and


ic are the exchange and cathodic current density respectively. The
ac and bc are measured from the experimental curves as shown in
Fig. 2. The cathodic overvoltage in the system under fretting cor-
rosion is primarily resulted from the overall reduction reaction on
the cathode (passive CpTi surrounding the fretting area) and oxi-
dation of anode (area under the fretting); this is represented by the
schematic shown in Fig. 7(b). In the present situation of the large
cathode and small anode (unequal area ratio), the net anodic cur-
rent density (e.g. ifretting ) will be accounted for the overvoltage. At
OCP, Ia = Ic , therefore ia Aa = ic Ac

ia Aa
Fig. 12. Variation in the ratio of worn out surface to the total surface area of CpTi ic = (6)
after various fretting durations at OCP in the Ringers solution. Ac

Using above relation, Eq. (5) becomes


sity (in the range of 295362 nA/cm2 ) (shown in Fig. 11). As the Aa
fretting progresses, the worn out area increases and corresponding  = ac + bc logia + bc log (7)
Ac
current density changes. The ratio of the worn out surface to the
total surface area after various fretting durations (5300 min) is ia and ic are the anodic and cathodic current densities respectively;
plotted in Fig. 12. The calculated fretting current densities using Ia and Ic are the anodic and cathodic current respectively; Aa and
Eq. (2) for the surface of different ra values (47443 nm) are Ac are the anode and cathode areas respectively. In present study,
5.239.2 nA/cm2 . The ifretting accounted for the fretting zone, are ia is substituted by the ifretting with corresponding fretted area (Aa ).
much larger than that observed during the static OCP measure- The ic is accounted for the area outside the fretting zone (Ac ). The
ments, iocp (current density at OCP). An almost constant current fretted area was measured after the various fretting durations and
obtained during fretting corrosion indicates an increase in the cur- plotted as Aa /Ac versus time, in Fig. 12. The Ac is considered equal
rent density since the fretting area is broadened with time. It is to the total surface area of the specimen (200 mm2 ) as Ac >> Aa .
this ifretting which causes the OCP drop (cathodic shift) during fret- Assuming the linear relationship between the Aa /Ac , and time (as
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 395

-100 -100

-200 (a) -150 (b)


Potential (mV/SCE)

Potential (mV/SCE)
-200 SIMUL
-300
EXP
-400 -250
SIMUL
-500 EXP -300
-600 -350
-700 -400
-800 -450
-900 -500
0 1000 2000 3000 0 1000 2000 3000
Time (s) Time (s)

-100 -100

-200
(c) -200 (d)

Potential (mV/SCE)
Potential (mV/SCE)

-300
-300 SIMUL
SIMUL
-400
EXP EXP
-400
-500
-500 -600
-600 -700

-700 -800
0 1000 2000 3000 0 1000 2000 3000
Time (s) Time (s)

Fig. 13. (ad) Experimental and simulated potential drop during the fretting corrosion of CpTi at OCP with the surface roughness (a) 43 nm, (b) 116 nm, (c) 177 nm and (d)
474 nm.

shown in Fig. 12), Eq. (7) can be related to the fretting time and be zone of the CpTi, inside and outside (smeared surface) of the wear
rewritten as track is shown in Fig. 14 (ai). The micro-cracks are also revealed
  at the fretted zone.
 = ac + bc logifretting + bc log (Aa /Ac + kt) (8)

where k is the constant of proportionality that depends on the ini- 3.4. Fretting corrosion at applied potential (+250 mV(SCE) )
tial surface condition such as roughness and materials type; t is
the fretting time (in seconds). The initial anodic to cathodic area Fretting corrosion experiments were performed on the CpTi
ratio (Aa /Ac at time t = 0) is a non-zero value which is obtained by specimens of various ra values (43474 nm) under the impressed
extrapolating the fretting area to the lower times (equal to Y-axis anodic potential of +250 mV(SCE) which is well within the passive
intercept). The Eq. (8) is used to simulate the cathodic potential zone (Fig. 2). The current versus time curves obtained during the
drop with the fretting time; the simulated potential drop curves fretting corrosion at +250 mV(SCE) is shown in Fig. 15(a). The anodic
due to the fretting corrosion are presented in Fig. 13(ad) for CpTi polarization experiments show the current of the order of 10 A
of different ra values. The ifretting , obtained from the fretting time- at +250 mV(SCE) for all the surfaces except for ra 43 nm. However,
current curves and Eq. (2) are 17.0, 5.2, 39.2, and 35.6 nA/cm2 this reduced to a much lower value 0.6 A under the poten-
for the surfaces with ra 43, 116, 177, and 474 nm respectively. The tiostatic condition (at constant potential +250 mV(SCE) ) after 1 h
simulated curves show a reasonable similarity with the experimen- (Fig. 15(a)). The current, in the beginning of the potentiostatic fret-
tal curves (Fig. 13(ad)). The point of difference, however, appears ting test, reached to the value of 10 A and dropped gradually
with respect to the experimental curve showing constant or a slight within 500 s of holding at the same potential. This is attributed
increase in the potential with fretting time, unlike the simulated to a larger increase in the resistance of the passive lm during
(dotted line in Fig. 13) curve. This is due to the fact that the oxide the potentiostatic hold compared to the anodic polarization scan.
formed during the fretting corrosion got cladded to the CpTi sur- The passive lm does not thicken due to the short residence time
face as brought out by the SEM investigation, shown in Fig. 14(a), (<1 s) during anodic polarization. After an hour of potentiostatically
and thereby reduced the anodic area. This resulted in an increase in holding the specimen, fretting ON enhanced the current (Itotal )
the potential towards more +ve value with the increase in fretting from 0.6 A to >2 A. This current value is though much smaller
time. Present efforts thus demonstrated, using a semi- empirical than that during the anodic polarization, yet approximately 3
relation, that the fretting current density attributed to the fretting times the value during steady state (potentiostatic hold without
area caused the potential drop. The simulated drop in OCP, thus, the fretting ON). It reduces further as fretting progresses up to
requires an initial contact area between the two counter surfaces 1800seconds. The current values during fretting initiation were
and the fretting current density. The simulation may be improved 2.33, 2.03, 2.3, and 2.05 A for CpTi with various ra from 43 to
further by addressing the complexities during fretting corrosion 474 nm which decreased to 0.43, 0.45, 0.98, and 0.49 A with the
such as more accurate treatise on the anodic area calculation with increase in fretting time. The corresponding maximum value of
the progress of time. The oxide and the wear debris on the fretting ifretting was the order of 10 mA/cm2 , while the minimum ifretting at
396 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

Fig. 14. (ai) (a) Inside the box shows a low magnication image of the fretted zone which is magnied to indicate the wear cracks and debris on the CpTi surface with ra
43 nm, (b) bulk elemental analysis of fretting zone, and elemental mapping of (c) Ti, (d) O, (e) Na, (f) Cl and (g) Al on fretting zone by SEM-EDAX, (h) smeared surface with
wear debris near fretting zone and (i) point analysis of wear debris.

the end of fretting cycle was 2.3 mA/cm2 . These are approximately ing the fretting corrosion indicate the exposure of a fresh surface
1000 times the value obtained during the anodic polarization at at an extremely localized region. These grew to the height ranging
+250 mV(SCE) . The decrease in fretting current density with time is between 38 A/cm2 which is lower than 10 but much larger than
possibly due to the cladding of fretting wear debris (TiO2 ) to the 1 A/cm2 .
CpTi that prevented the exposure of fresh surface to the Ringers Fig. 15(b) illustrates the variation in charge density resulting
solution. The current transients, frequently found to appear dur- from the mass loss of CpTi with different ra at an applied potential
B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398 397

Fig. 15. (ac) Variation in (a) current density, (b) corresponding charge density during the fretting corrosion at an applied potential of +250 mV(SCE) (>Ecorr ) and (c) charge
density during the fretting corrosion of CpTi (with ra 43474 nm) at OCP in Ringers solution.

of +250 mV(SCE) . An initial large slope up to 1000seconds delin- tion. It may be expected that after a certain fretting duration, the
eates increasing corrosion of the freshly exposed CpTi followed surface undulation reaches to the similar level irrespective of an ini-
by a smaller slope (>1000 s) owing to the oxide thickening and tial roughness. The rate of increase in the charge density, thereafter,
subsequent sluggish corrosion kinetics. The charge consumed is may become same for the specimens with all roughnesses.
used for passive lm thickening rather than for corrosion or mass
loss. After the fretting ON at >3000 s, slope of the curve is further
4. Conclusions
enhanced due to the accelerated electrochemical activity. Initial
segment, within the fretting ON zone (up to 3800 s), demonstrates
This study brought out the role of surface roughness on corro-
a linear increase in the charge density associated with the abra-
sion and fretting corrosion resistance of the CpTi in the Ringers
sion assisted corrosion. A relatively slower growth of the charge
solution. The following inferences from the study may be useful
density from 3800seconds up to the fretting end could be due to
during the fabrication or surface engineering of the titanium based
the cladding of oxide and debris which alleviated the surface area
implants:
of CpTi for fretting and thus reduced the corrosion. At OCP (with-
out an applied potential), the charge density demonstrated a linear
increase irrespective of the fretting ON or OFF cycle, as illus- 1. The anodic polarization and EIS study revealed better corrosion
trated in Fig. 15(c). The magnitude by which the charge density resistance of CpTi surface with ra 43 nm than that beyond 43 nm
is increased due to fretting ON is signicantly small, as marked by in Ringers solution. The surface energy of the CpTi with varied
I and II in Fig. 15(c), compared to that at an applied potential. Since roughness corresponds well with their corrosion resistance.
the increase in corrosion activity due to the fretting at OCP is small, 2. Severe corrosion damage is observed, when fretting is performed
the resultant variation in the slope of the charge density-time curve at an applied anodic potential, as indicated by the ifretting which
is insignicant. is several order higher than during the potentiodynamic polar-
The charge density curve depicts a larger mass loss for the rough ization or OCP.
surface compared to the smooth. Overall, charge density accumu- 3. A drop in potential simulated, using fretting corrosion cur-
lated for ra 43 and 116 nm is much smaller compared to that for rent density obtained from the experimental results, has shown
ra 474 and 177 nm, with and without applied potential. The higher excellent tting. The knowledge of ifretting , in addition to an ini-
charge density and the mass loss associated with the greater sur- tial contact area is, therefore, necessary to simulate the potential
face roughness could be due to the instability of passive lm at drop to predict the surface damage with high accuracy.
sharp peaks, which under repeated fretting suffers severe destruc- 4. The highest drop in potential due to the fretting corrosion of CpTi
with ra 43 nm and corresponding worn out surface area suggests
398 B. Sivakumar et al. / Applied Surface Science 401 (2017) 385398

a larger surface damage; this, however, needs to be studied for [19] Y. Li, Y.F. Cheng, Effect of surface nishing on early-stage corrosion of a
the longer fretting durations. carbon steel studied by electrochemical and atomic force microscope
characterizations, Appl. Surf. Sci. 366 (2016) 95103.
[20] M. Es-Souni, M. Es-Souni, H. Fischer-Brandies, On the properties of two binary
Acknowledgements NiTi shape memory alloys. Effects of surface nish on the corrosion behaviour
and in vitro biocompatibility, Biomaterials 23 (2002) 28872894.
[21] D. Landolt, S. Mischler, M. Stemp, Electrochemical methods in tribocorrosion:
Bose Sivakumar expresses his thanks to CSIR, New Delhi for pro- a critical appraisal, Electrochim. Acta 46 (2001) 39133929.
viding senior research fellowship. This work is partly supported by [22] C. Colombie, Y. Berthier, A. Floquet, L. Vincent, M. Godet, Fretting:
the project OLP-0174 at CSIR-NML. Mr. Y.N. Singhbabu, AcSIR PhD load-carrying capacity of wear debris, ASME J. Tribol. 106 (1984) 185194.
[23] Y. Fu, J. Wei, A.W. Batchelor, Some considerations on the mitigation of fretting
Scholar, is acknowledged for his kind help in XPS analysis. Authors damage by the application of surface-modication technologies, J. Mater.
sincerely thank the Director, CSIR-NML for his continuous support Process. Technol. 99 (2000) 231245.
and permission to publish this research. [24] A. Panagiotidou, J. Meswania, J. Hua, S. Muirhead-Allwood, A. Hart, G. Blunn,
Enhanced wear and corrosion in modular tapers in total hip replacement is
associated with the contact area and surface topography, J. Orthop. Res. 31
References (2013) 20322039.
[25] P. Ponthiaux, F. Wenger, D. Drees, J.P. Celis, Electrochemical techniques for
[1] S. Barril, S. Mischler, D. Landolt, Electrochemical effects on the fretting studying tribocorrosion processes, Wear 256 (2004) 459468.
corrosion behavior of Ti6Al4V in 0.9% sodium chloride solution, Wear 259 [26] A.C. Vieira, L.A. Rocha, N. Papageorgiou, S. Mischler, Mechanical and
(2005) 282291. electrochemical deterioration mechanisms in the tribocorrosion of Al alloys
[2] L. Duisabeau, P. Combrade, B. Forest, Environmental effect on fretting of in NaCl and in NaNO3 solutions, Corros. Sci. 54 (2012) 2635.
metallic materials for orthopaedic implants, Wear 256 (2004) 805816. [27] J.P. Celis, P. Ponthiaux, F. Wenger, Tribo-corrosion of materials: interplay
[3] N. Diomidis, S. Mischler, N.S. More, M. Roy, S.N. Paul, Fretting-corrosion between chemical, electrochemical and mechanical reactivity of the surfaces,
behavior of titanium alloys in simulated synovial uid, Wear 271 (2011) Wear 261 (2006) 939946.
10931102. [28] D. Mareci, G. Ungureanu, D.M. Aelenei, J.C. Mirza Rosca, Electrochemical
[4] J. Komotori, N. Hisamori, Y. Ohmori, The corrosion/wear mechanisms of characteristics of titanium based biomaterials in articial saliva, Mater.
Ti-6Al-4V alloy for different scratching rates, Wear 263 (2007) 412418. Corros. 58 (2007) 848856.
[5] S. Kumar, B. Sivakumar, T.S.N. Sankara Narayanan, S.G. Sundara Raman, S.K. [29] M.V. Popa, I. Demetrescu, E. Vasilescu, P. Drob, A. Santana Lopez, J.
Seshadri, Fretting-corrosion mapping of CP-Ti in Ringers solution, Wear 268 Mirza-Rosca, C. Vasilescu, D. Ionita, Corrosion susceptibility of implant
(2010) 15371541. materials Ti-5Al-4V and Ti-6Al-4Fe in articial extra-cellular uids,
[6] S. Kumar, T.S.N. Sankara Narayanan, S.G. Sundara Raman, S.K. Seshadri, Electrochim. Acta 49 (2004) 21132121.
Fretting corrosion behaviour of thermally oxidized CP-Ti in Ringers solution, [30] S.L. Assis, S. Wolynec, I. Costa, Corrosion characterization of titanium alloys by
Corros. Sci. 52 (2010) 711721. electrochemical techniques, Electrochim. Acta 5 (2006) 18151819.
[7] L. Benea, E. Danaila, P. Ponthiaux, Effect of titania anodic formation and [31] J. Pan, D. Thierry, C. Leygraf, Electrochemical impedance spectroscopy study
hydroxyapatite electrodeposition on electrochemical behavior of Ti-6Al-4V of the passive oxide lm on titanium for implant application, Electrochim.
alloy under fretting conditions for biomedical applications, Corros. Sci. 91 Acta 41 (1996) 11431153.
(2015) 262271. [32] N. Ibris, J.C.M. Rosca, EIS study of Ti and its alloys in biological media, J.
[8] R.K. Alla, K. Ginjupalli, N. Upadhya, M. Shammas, R.K. Ravi, R. Sekhar, Surface Electroanal. Chem. 526 (2002) 5362.
Roughness of Implants: a review, Trends Biomater. Artif. Organs 25 (2011) [33] C.E.B. Marino, L.H. Mascaro, EIS characterization of a Ti-dental implant in
112118. articial saliva media: dissolution process of the oxide barrier, J. Electroanal.
[9] T. Albrektsson, T. Berglundh, J. Lindhe, Osseointegration: historic background Chem. 568 (2004) 115120.
and current concepts, in: Clinical Periodontology and Implant Dentistry, 4th [34] N.P. Hunt, S.J. Cunningham, C.G. Golden, M. Sheriff, An investigation into
ed., Blackwell, Munksgaard, Oxford, 2003, pp. 809820. effects of polishing on surface hardness and corrosion of orthodontic arch
[10] L. Le Guehennec, A. Soueidan, P. Layrolle, Y. Amouriq, Surface treatments of wires, Angle Orthod. 69 (1999) 433440.
titanium dental implants for rapid osseointegration, Dent. Mater. 23 (2007) [35] D. Khang, J. Lu, C. Yao, The role of nanometer and sub-micron surface features
844854. on vascular and bone cell adhesion on titanium, Biomaterials 29 (2008)
[11] N. Mirhosseini, P.L. Crouse, M.J.J. Schmidth, L. Li, D. Garrod, Laser surface 970983.
micro-texturing of Ti-6Al-4V substrates for improved cell integration, Appl. [36] A. Fattah-alhosseini, M.A. Golozar, A. Saatchi, K. Raeissi, Effect of solution
Surf. Sci. 253 (2007) 77387743. concentration on semiconducting properties of passive lms formed on
[12] K. Sasaki, G.T. Burstein, The generation of surface roughness during slurry austenitic stainless steels, Corros. Sci. 52 (2010) 205209.
erosion-corrosion and its effect on the pitting potential, Corros. Sci. 38 (1996) [37] C.A. Della Rovere, J.H. Alano, R. Silva, P.A.P. Nascente, J. Otubo, S.E. Kuri,
21112120. Characterization of passive lms on shape memory stainless steels, Corros.
[13] P.E. Manning, D.J. Duquette, W.F. Savage, The effect of test method and Sci. 57 (2012) 154161.
surface condition on pitting potential of single and duplex phase 304L [38] J.-C. Dupin, D. Gonbeau, P. Vinatier, A. Levasseur, Systematic XPS studies of
stainless steel, Corros. Sci. 35 (1979) 151157. metal oxides, hydroxides and peroxides, Phys. Chem. Chem. Phys. 2 (2000)
[14] G.E. Coates, Effect of some surface treatments on corrosion of stainless steel, 13191324.
Mater. Perf. 29 (1990) 6167. [39] E. McCafferty, J.P. Wightman, An X-ray photoelectron spectroscopy sputter
[15] R. Lage, P. Moller, H.E. Fallesen, The effect of surface treatment and prole study of the native air-formed oxide lm on titanium, Appl. Surf. Sci.
topography on corrosion behavior of EN 1.4404 stainless steel, Mater. Corros. 143 (1999) 92100.
24 (2014) 10601067. [40] E. McCafferty, Introduction to Corrosion Science, Springer, New York,
[16] T. Hong, M. Nagumo, Effect of surface roughness on early stages of pitting Dordrecht, Heidelberg, London, 2010.
corrosion of Type 301 stainless steel, Corros. Sci. 39 (1997) 16651672. [41] E. Protopopoff, P. Marcus, Potential Versus pH (Pourbaix) Diagrams,
[17] M.B. Leban, C. Mikyska, T. Kosec, B. Markoli, J. Kovac, The effect of surface Corrosion: Fundamentals, Testing, and Protection, Vol 13A, ASM Handbook,
roughness on the corrosion properties of type AISI 304 stainless steel in ASM International, 2003, pp. 1730.
diluted NaCl and urban rain solution, J. Mater. Eng. Perf. 23 (2014) 16951702. [42] D.C. Sutton, G. Limbert, D. Stewart, R.J.K. Wood, A functional form for wear
[18] S.K. Kim, I.J. Park, D.Y. Lee, J.G. Kim, Inuence of surface roughness on the depth of a ball and a at surface, Tribol. Lett. 53 (2014) 173179.
electrochemical behavior of carbon steel, J. Appl. Electrochem. 43 (2013) [43] A.F. Bower, Applied Mechanics of Solids, CRC press, Taylor & Francis Group,
507514. Boca Raton, London, New York, 2009, pp. 6987.

Você também pode gostar