comparison of fretting corrosion behaviour of

Upload: vivianita240226326

Post on 02-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    1/7

    Comparison of fretting corrosion behaviour ofTi6Al4V alloy and CP-Ti in Ringers solution

    B. Sivakumar, S. Kumar and T. S. N. Sankara Narayanan*

    The fretting corrosion behaviour of Ti6Al4V alloy in Ringers solution was evaluated and

    compared with that of commercially pure titanium (CP-Ti). Free corrosion potential, morphology of

    the fretted zone, extent of formation and accumulation of debris and wear volume were used as

    parameters of evaluation. Both Ti6Al4V alloy and CP-Ti behave similarly in terms of change in

    free corrosion potential as a function of time, morphological features and wear mechanism. Ti

    6Al4V alloy, however, exhibits an increase in corrosion susceptibility, decrease in tendency for

    repassivation, higher amount of formation and accumulation of debris and an increase in wear

    volume compared with CP-Ti. The study points out the importance of material selection for

    implants that would encounter fretting corrosion.

    Keywords: Tribocorrosion, Fretting corrosion, Ti and its alloys, Implant, Joint prostheses, Repassivation

    Introduction

    Titanium and its alloys are widely used as orthopaedicand dental implants due to their low density, better

    mechanical properties, very high strength/weight ratio(specific strength), excellent corrosion resistance and

    biocompatibility.15 Among the various types of Tialloys, Ti6Al4V alloy has been the choice in manyinstances because its mechanical properties and corrosion

    resistance are ideal for implant applications. Studies onthe corrosion and biocompatibility aspects of Ti and itsalloys performed in vitro proved that the passive oxide

    layer is stable and offers excellent corrosion protectionand biocompatibility.15 Implant retrieval analysis, how-

    ever, reveals discolouration of the implant and accumula-tion of metal ions on tissues beside the implant.6 Theinferior mechanical properties of the naturally formed

    passive oxide layer that could be disrupted at very lowshear stresses, even by rubbing against soft tissues, areconsidered responsible for such an occurrence.7 Owing to

    the inherent property of titanium and its alloys, thepassive oxide layer could subsequently reform upon

    reaction with the local environment. However, implantretrieval analysis confirms that the capability of restora-tion of the damagedpassive film is not instantaneous, as it

    is generally believed.

    Fretting corrosion is the deterioration of a materialthat occurs at the interface of two contacting surfaces

    due to small oscillatory movements in the presence of acorrosive medium. Manufacturing of implant materials,though involves a component geometry specific locking

    mechanism, micromotion do occur,8 which enables thepenetration of the body fluid into this junction and

    facilitates mechanically assisted crevice and frettingcorrosion.9,10 The modular interfaces of total joint

    prosthesis, mainly at the fixation of the implant stem

    bone or cement, are subjected to micromotion (,100

    mm) that could result in fretting corrosion.11,12 Thefretting corrosion behaviour of untreated and surface

    modified titanium and its alloys was studied by many

    researchers.1320 These studies confirm the following

    observations:(i) removal of the passive oxide layer induced by

    fretting

    (ii) formation and entrapment of debris at thefretted zone, though most of them are pushed

    away towards edges

    (iii) increase in wear volume if the debris possesses

    an abrasive character

    (iv) delay in repassivation after the fretting motion

    is ceased.

    The fretting corrosion behaviour of Ti and its alloyscould be different in terms of the nature of the passive

    film, susceptibility for corrosion upon removal of the

    passive film, hardness of the alloy, extent of fretting

    wear, abrasive nature of the oxide debris, rate ofcorrosion and leaching of alloying elements. Such a

    comparison will be of much help to choose the right type

    of material for implants that would encounter fretting

    corrosion. In this perspective, the present work aims toevaluate the fretting corrosion behaviour of Ti6Al4V

    alloy in Ringers solution and compare it with that of

    commercially pure titanium (CP-Ti).

    Experimental

    Commercially pure titanium (grade 2) [with the chemi-

    cal composition of Ti0?01N0?03C0?01H0?20Fe

    0?18O (wt-%)] and Ti6Al4V alloy (grade 5) [with thechemical composition of Ti0?02N0?03C0?011H

    0?22Fe0?16O6?12Al3?93V (wt-%)] discs of 20 mm in

    Metallurgical Laboratory, Madras Centre, CSIR Complex, Taramani,

    Chennai 600 113, India

    *Corresponding author, email [email protected]

    ;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;

    2011 W. S. Maney & Son LtdReceived 3 September 2011; accepted 12 October 2011

    15 8 Tribology 2011 VOL 5 NO 4 DOI 10.1179/1751584X11Y. 0000000020

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    2/7

    diameter and 2 mm in thickness were used as substrate

    materials. They were mechanically polished using

    various grades of SiC paper followed by 0?3 mmdiamond paste to a mirror finish, rinsed with deionised

    water and dried using a stream of compressed air.

    Fretting corrosion experiments were performed using a

    fretting corrosion test assembly. The details of the test

    assembly have already been discussed in our earlier

    papers.1720 A ball on flat contact configuration that

    involves a 8 mm w alumina ball (finish, G 10 grade;hardness, 1365 HV) moving against the stationary CP-

    Ti/Ti6Al4V alloy flat was chosen so that large contact

    stresses could be achieved under very low loads.

    Normal loads of 3 and 10 N, oscillating frequencies of

    5 and 10 Hz and linear peak to peak displacement

    amplitude of 180 mm were used as the fretting corrosiontest parameters. The Hertzian contact pressures for the

    loads used (3 and 10 N) will be ,500 and 1200 MPa.

    The tests were performed for 18 000 (5 Hz) and 36 000

    (10 Hz) fretting cycles. The test parameters employed in

    this study imply a gross slip condition. Ringers solution,

    having a chemical composition (in g L21) of 9NaCl

    0?24CaCl20?43KCl0?2NaHCO3 (pH 7?8) at 310 K,

    was used as the electrolyte solution. The CP-Ti/Ti

    6Al4V alloy discs subjected to fretting corrosion

    formed the working electrode, while a saturated calomel

    electrode (SCE) and a graphite rod served as reference

    and auxiliary electrodes respectively. These electrodes

    were placed in the fretting corrosion cell in such a way

    that only 2 cm2 area of the working electrode wasexposed to the Ringers solution. The alumina ballCP-

    Ti/Ti6Al4V alloy flat contact was arranged in such a

    way that they were totally immersed in the Ringers

    solution. The fretting corrosion cell was connected to

    a potentiostat/galvanostat/frequency response analyser

    from ACM Instruments (model Gill AC) to measure the

    free corrosion potential (FCP) of CP-Ti/Ti6Al4V

    alloy as a function of time. Before the onset of fretting,

    CP-Ti/Ti6Al4V alloy was allowed to stabilise for 1 h

    in Ringers solution. The change in FCP of CP-Ti/Ti

    6Al4V alloy was monitored as a function of time. The

    FCP measurement was repeated at least three times to

    ensure reproducibility of the test results. The morpho-logical features of the fretted zone were assessed using

    SEM. Energy dispersive X-ray (EDX) analysis was

    performed at selected regions of the fretted zone to

    identify their chemical nature. The three-dimensional

    (3D) profile of the fretted zone was assessed using anultrasonic based non-destructive testing device.

    Results and discussion

    Fretting corrosion behaviour of Ti6Al4V alloyFree corrosion potential measurement of Ti6Al4V alloy

    The FCP is a qualitative indicator of the corrosion

    regime (active or passive), in which a metal resides, andit has been used to evaluate the performance of Ti and

    its alloys under fretting corrosion conditions.2024 Thechange in FCP of Ti6Al4V alloy recorded before the

    onset of fretting, with the onset of fretting, during

    fretting and after the fretting motion is ceased, is shownin Fig. 1a. The FCP is a mixed potential, reflecting the

    state of the unworn material and those in the frettingwear track. Before the onset of fretting, the FCP of Ti

    6Al4V alloy exhibits an anodic shift, suggestingthickening of the passive film during the initial

    stabilisation period of 1 h. With the onset of fretting, a

    sudden drop in FCP (cathodic shift) (Fig. 1a) anda surge in anodic current (Fig. 1b) are observed. A

    similar observation was also made earlier by many

    researchers during tribocorrosion of bare and surfacemodified Ti, Ti6Al4V alloy and stainless steel in many

    environments.13,14,1723 Komotoriet al.24 have observed

    these changes when Ti6Al4V alloy is scratched by a

    sapphire ball in Ringers solution. According toPonthiaux et al.,25 the FCP of titanium becomes muchcloser to the freshly ground material in the electrolyte

    during corrosion wear. The extent of cathodic shift in

    FCP and the surge in anodic current observed in thepresent study indicates removal of the passive oxide

    layer and increase in corrosion susceptibility of Ti6Al4V alloy in Ringers solution. The increase in applied

    load from 3 to 10 N results in a higher cathodic shift inFCP (Fig. 1a). This is due to the effective removal of the

    passive film that results in a larger active fretted area.

    Contuet al.26 have also observed a similar effect duringthe mechanical abrasion of Ti and Ti6Al4V alloy in

    inorganic buffer. The increase in frequency, from 5 to10 Hz, seems to have a relatively lesser influence than

    those caused by the load (Fig. 1a).

    1 Change in a FCP and b anodic current of Ti6Al4V alloy measured as function of time (fixed condition: amplitude,

    180 mm; variables: load, frequency and fretting cycles)

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    Tribology 2011 VOL 5 NO 4 15 9

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    3/7

    During fretting, some fluctuations in the FCP of Ti

    6Al4V alloy are observed following the periodicremoval (depassivation) and growth (repassivation) of

    the passive oxide layer in the fretted zone, suggesting the

    existence of a dynamic equilibrium between depassiva-

    tion and repassivation phenomena.14,22 The averagevalues of fluctuation in FCP for a load of 3 N are 902

    and 1122 mV at 5 and 10 Hz respectively. However,

    when the load is increased from 3 to 10 N, the

    corresponding values become 662 and 762 mV.

    The decrease in the extent of fluctuations with the

    increase in load indicates the decrease in tendency of

    the alloy to repassivate. This can be attributed to theincrease when the load is increased from 3 to 10 N. After

    the fretting motion is ceased, the FCP of Ti6Al4V

    alloy exhibits an anodic shift, suggesting the occurrenceof repassivation. This is due to the rapid regeneration of

    TiO2 layer in the active areas of the fretted zone

    following the reaction of the fresh Ti metal ions with the

    dissolved oxygen available in Ringers solution. Asimilar behaviour was observed earlier by Komotori

    et al.24 and Ponthiauxet al.25 during the repassivation of

    Ti and Ti6Al4V alloy. During repassivation, two

    important factors, such as the ability of the material to

    return to the initial steady state potential and the time

    required for such an occurrence, should be considered.Ideally, the potential should reach the initial steady state

    before the onset of fretting. After the fretting motion is

    ceased, Ti6Al4V alloy attained its initial steady state

    potential for a load of 3 N at 5 and 10 Hz. However,

    when the load is increased to 10 N, though the FCP

    reaches the initial steady state potential, the time taken

    for this occurrence becomes relatively higher. The

    increase in contact area of the fretted zone and the

    2 Morphology of Ti6Al4V alloy after subjecting it to fretting corrosion (conditions: amplitude, 180 mm; frequency,

    10 Hz; load, 10 N; fretting cycles, 36 000): a fretted zone (circled area) and surrounding areas (fretting direction is indi-

    cated by doubled sided arrow mark); b central region of fretted zone; c debris collected at edges

    3 Change in FCP of CP-Ti and Ti6Al4V alloy measured as function of time:a amplitude, 180 mm; load, 3 N; frequency, 5 Hz;

    number of fretting cycles, 18 000;bamplitude, 180 mm; load, 10 N; frequency, 5 Hz; number of fretting cycles, 18 000

    4 Rate of change in FCP of CP-Ti and Ti6Al4V alloymeasured during repassivation (after fretting motion is

    ceased)

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    16 0 Tribology 2011 VOL 5 NO 4

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    4/7

    extent of damage at 10 N could be considered respon-sible for this behaviour.

    Surface morphology of fretted zone of Ti6Al4V alloy

    The morphology of Ti6Al4V alloy after frettingcorrosion (conditions: amplitude, 180 mm; frequency,10 Hz; load, 10 N; fretting cycles, 36 000) is shown inFig. 2.

    The fretted zone (circled region) has experiencedsevere damage, whereas the surrounding areas of thefretted zone are relatively smooth, in which wear debrisis smeared all around (Fig. 2a). The central region of thefretted zone reveals severe damage due to the extensiveshear deformation and the ploughing action of the

    alumina ball, suggesting the involvement of adhesivegalling as the predominant wear mechanism (Fig. 2b).Microwelding of surface asperities occurs during the

    initial stages, whereas the asperities get sheared andplucked away in the subsequent stages. Redeposition ofthe removed material, confirmed by the presence ofdebris within the fretted zone (Fig. 2b), enables anincrease in roughness and further accelerates the wearrate. The debris collected at the edges of the fretted zoneis shown in Fig. 2c. The surface of the Al2O3 ballcounterface reveals the transfer of material from Ti

    6Al4V alloy, which confirms the occurrence of adhesivegalling.

    Comparison of fretting corrosion behaviour ofCP-Ti and Ti6Al4V alloyThe similarity in shape of the FCPtime curves of CP-Ti

    and Ti6Al4V alloy suggests the occurrence of similarphenomena during fretting corrosion of these materials(Fig. 3).

    5 Comparison of morphologies of a, c, e, g CP-Ti and b, d, f, h Ti6Al4V alloy after subjecting them to fretting corro-

    sion (conditions: amplitude, 180mm; load, 10 N; frequency, 5 Hz; fretting cycles, 18 000): a, b entire fretted zone

    (circled region; fretting direction is indicated by double sided arrow mark); c, d central region;

    e, f edge region; g, h debris at edges

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    Tribology 2011 VOL 5 NO 4 16 1

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    5/7

    In spite of the similarity in trend, some noticeable

    differences could be observed. Compared with CP-Ti,

    Ti6Al4V alloy exhibits a higher cathodic shift in FCP

    with onset fretting, few fluctuations during fretting and a

    decrease in the rate of anodic shift in FCP after thefretting motion is ceased. These effects are well

    pronounced at 10 N (Fig. 3b). The higher cathodic shift

    in FCP signifies the increase in susceptibility of Ti6Al

    4V alloy for corrosion in Ringers solution. The fewfluctuations in FCP indicate the decrease in tendency of

    Ti6Al4V alloy to repassivate. Contu et al.26 have also

    reported that CP-Ti displays a better tendency for

    repassivation than Ti6Al4V alloy in inorganic buffer

    solution at pH 4?0 a n d 7?0. The delay in cathodic

    reaction kinetics can be considered responsible for the

    poor tendency for repassivation exhibited by Ti6Al4V

    alloy compared with CP-Ti.27

    The rate of change in FCP after the fretting motion is

    ceased confirms the decrease in ability of Ti6Al4V

    alloy to revert to the initial steady state (Fig. 4). The

    time to reach a threshold value of 2550 mV(SCE) is

    117 s for CP-Ti, whereas Ti6Al4V alloy under similarexperimental conditions (at 10 N/5 Hz) attains this

    threshold only at 672 s.

    A comparison of the morphologies of CP-Ti and Ti

    6Al4V alloy after subjecting them to fretting corrosion

    (conditions: amplitude, 180 mm; load, 10 N; frequency,

    5 Hz; fretting cycles, 18 000) is shown in Fig. 5. The

    fretted zone (circled region) has experienced severe

    damage due to the extensive shear deformation and

    ploughing action of the alumina ball (Fig. 5a and b),

    suggesting the involvement of adhesive galling as the

    predominant wear mechanism in both cases. However,the central (Fig. 5cand d) and edge regions (Fig. 5eand

    f) of the fretted zone reveal that the amount of

    formation of debris and their entrapment is relatively

    high for Ti6Al4V alloy. In addition, the extent of

    accumulation of debris in the edge regions of the fretted

    zone is relatively high for Ti6Al4V alloy (Fig. 5gand

    h).

    The 3D profile of the fretted zone of CP-Ti and Ti

    6Al4V alloy, after subjecting them to fretting corrosion

    (conditions: amplitude, 180 mm; load, 3 N; frequency,

    5 Hz; fretting cycles, 18 000), is shown in Fig. 6a and b

    respectively.

    It is evident from Fig. 6 that the wear volume is higherfor Ti6Al4V alloy than for CP-Ti under similar

    conditions. Masmoudi et al.28 have pointed out that

    6 Three-dimensional profile of fretted zone of a CP-Ti and b Ti6Al4V alloy after fretting corrosion (conditions: ampli-

    tude, 180mm; load, 3 N; frequency, 5 Hz; fretting cycles, 18 000) (X and Y axes: values are in mm; Z axis: values are

    in 61023 mm)

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    16 2 Tribology 2011 VOL 5 NO 4

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    6/7

    the wear rate of nitric acid passivated Ti6Al4V alloy is

    higher than that of CP-Ti in Ringers solution.

    According to them, the lower thickness of the oxidefilm and its inferior resistance to corrosive medium are

    responsible for the higher wear rate of Ti6Al4V

    alloy.28 This observation is also supported by other

    researchers.27,29,30 Contuet al.26 have reported that with

    respect to corrosion, both CP-Ti and Ti6Al4V alloy

    exhibit similar behaviours during mechanical abrasion

    in inorganic buffer, since oxidation of titanium is the

    major reaction. The tendency for repassivation, how-

    ever, is relatively higher for CP-Ti than for Ti6Al4V

    alloy.26 Martinet al.31 have suggested that the hardness

    of wear debris becomes the controlling factor in

    determining the performance of Ti6Al4V alloy under

    tribocorrosion conditions. The nanohardness of Ti6Al

    4V alloy is ,5?1 GPa, whereas the hardness of its debris

    layer is reported to be nearly double.32

    The decrease in ability of Ti6Al4V alloy to revert to

    its initial steady state (Fig. 4) could be correlated to the

    extent of formation and entrapment of debris at the

    fretted zone, the hard and abrasive nature of the debris

    and the increase in wear volume. The morphological

    features (Fig. 5) of the fretted zone of Ti6Al4V alloy

    confirm the generation of higher quantities of debris and

    their entrapment in the fretted zone. The EDX analysis

    performed in the regions marked as % of the fretted

    zone of CP-Ti and Ti6Al4V alloy reveals their

    chemical nature (Fig. 7). For CP-Ti, this region contains

    67?83 at-% of Ti, 25?96 at-% of oxygen and 6?21 at-% ofAl (Fig. 7a), which indicates that it is predominantly

    oxides of Ti. The presence of Al could have originated

    from the alumina ball used as the counterface. For Ti

    6Al4V alloy, this region contains 41?53 at-% of Ti,

    47?09 at-% of oxygen, 8?93 at-% of Al, 1?44 at-% of V

    and 1?01 at-% of Cl (Fig. 7b).

    The higher oxygen content indicates that the extent of

    oxidation of the fretted zone is relatively higher for Ti

    6Al4V alloy compared with that of CP-Ti. The

    presence of Al and V, with a corresponding decrease

    in Ti, supports the formation of oxides of Al and V in

    the fretted zone of Ti6Al4V alloy. The hard and

    abrasive nature of the oxides of Al would have increasedthe wear rate/wear volume of Ti6Al4V alloy, which is

    confirmed by the 3D profile of the fretted zone (Fig. 6).

    The higher cathodic shift in FCP of Ti6Al4V alloy

    with the onset of fretting, the decrease in tendency of the

    alloy to repassivate during fretting and the decrease inability of the alloy to revert to the initial steady state

    after the fretting motion is ceased assume significance.

    The increase in corrosion susceptibility with the removal

    of the passive layer induced by fretting, the poor

    tendency for repassivation during fretting and the delay

    in reaching the initial steady state potential would

    induce leaching of Al and V ions, which could cause

    long term health problems like Alzheimer disease and

    neuropathy. Osteolysis, adverse tissue reactions, kidney

    lesion, cytotoxicity, hypersensitivity and carcinogenesis

    have been reported to be associated with V and Al

    ions.5,3335 Vanadium may elicit local or especially

    systemic reactions or inhibit cellular proliferation.

    Aluminium may be associated with osteomalacia,

    pulmonary granulomatosis and neurotoxicity.34,35 The

    accumulation of wear debris may produce an adverse

    cellular response, leading to inflammation, release of

    damaging enzymes, osteolysis, infection, implant loosen-

    ing and pain.36,37

    Conclusion

    The study on the fretting corrosion behaviour of Ti

    6Al4V alloy in Ringers solution and comparison of its

    behaviour with that of CP-Ti lead to the following

    conclusions.

    1. Both CP-Ti and Ti6Al4V alloy exhibit a similartrend of cathodic shift in FCP with the onset of fretting,

    fluctuations in FCP during fretting and anodic shift in

    FCP after the fretting motion is ceased. Adhesive galling

    is the predominant wear mechanism when they are

    fretted against the alumina ball.

    2. The fretting corrosion behaviour of Ti6Al4V

    alloy differs from that of CP-Ti in terms of: increase in

    susceptibility for corrosion upon removal of the passive

    oxide layer with the onset of fretting, decrease in

    tendency to repassivate during fretting, decrease in

    ability to revert to the initial steady state potential after

    the fretting motion is ceased, higher amount of

    formation and entrapment of debris at the fretted zoneand accumulation of the same at the edges, increased

    extent of oxidation leading to the formation of oxides of

    7 Analysis (EDX) performed on marked region: region marked as % in Fig. 5c and d of fretted zone of a CP-Ti and b Ti

    6Al4V alloy (conditions: as described in Fig. 5)

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    Tribology 2011 VOL 5 NO 4 16 3

  • 8/11/2019 Comparison of Fretting Corrosion Behaviour Of

    7/7

    Al and V besides Ti and increase in wear volume due tothe abrasive nature of aluminium oxide.

    3. The difference in performance of Ti6Al4V alloyand CP-Ti points out that the choice of materials for

    implants that would encounter fretting corrosion shouldbe made only after a thorough evaluation.

    Acknowledgement

    The authors express their sincere thanks to Dr S.Srikanth, Director, National Metallurgical Laboratory,Jamshedpur, for his constant encouragement and sup-port and for his permission to publish this paper.

    References1. D. F. Williams: in Biocompatibility of clinical implant materials,

    (ed. D. F. Williams), Vol. II, 944; 1981, Boca Raton, FL, CRC

    Press.

    2. P. Kovacs and J. A. Davidson: in Medical applications of titanium

    and its alloys: the materials and biological issues, (ed. S. A. Brown

    and J. E. Lemons), ASTM STP 1272, 63; 1996, Philadelphia, PA,

    American Society of Testing and Materials.

    3. M. Long and H. J. Rack: Biomaterials, 1998, 19, 16211639.

    4. J. A. Hunt and M. Shoichet:Curr. Opin. Solid State Mater. Sci. ,2001, 5, 161162.

    5. M. Geetha, A. K. Singh, R. Asokamani and A. K. Gogia: Prog.

    Mater. Sci., 2009, 54, 397425.

    6. Y. Mu, T. Kobayashi, M. Sumita, A. Yamamoto and T. Hanawa:

    J. Mater. Sci. Mater. Med., 2002, 13, 583588.

    7. P. A. Lilley, P. S. Walker and G. W. Blunn: Proc. 4th World

    Biomaterials Cong., Berlin, Germany, April 1992, 227230.

    8. J. J. Jacobs, J. L. Gilbert and R. M. Urban: J. Bone Joint Surg.,

    1998, 80, 268282.

    9. J. L. Gilbert and J. J. Jacobs: in Modularity of orthopedic

    implants, (ed. J. E. Parr et al.), ASTM STP 1301, 4559; 1997,

    Philadelphia, PA, American Society of Testing and Materials.

    10. J. J. Jacobs, J. L. Gilbert and R. M. Urban: in Advances in

    orthopaedic surgery, (ed. R. N. Stauffer), Vol. 2, 279319; 1994, St

    Louis, MO, Mosby.

    11. D. W. Hoeppner and V. Chandrasekaran: Wear, 1994, 173, 189

    197.12. M. Windler and R. Klabunde: in Titanium in medicine, (ed. D. M.

    Brunette et al.), Vol. 1, 703746; 2001, Berlin, Berlin, Springer-Verlag.

    13. S. Barril, S. Mischler and D. Landolt:Wear, 2005, 259, 282291.

    14. B. Tang, P. Q. Wu, A. L. Fan, L. Qin, H. J. Hu and J.-P. Celis:

    Adv. Eng. Mater., 2005, 7, 232238.

    15. S. Hiromoto and S. Mischler:Wear, 2006, 261, 10021011.

    16. A. C. Vieira, A. R. Ribeiro, L. A. Rocha and J. P. Celis: Wear,

    2006, 261, 9941001.

    17. S. Kumar, T. S. N. Sankara Narayanan, S. Ganesh Sundara

    Raman and S. K. Seshadri: Corros. Sci., 2010, 52, 711721.

    18. S. Kumar, B. Sivakumar, T. S. N. Sankara Narayanan, S. Ganesh

    Sundara Raman and S. K. Seshadri: Wear, 2010, 268, 15371541.

    19. S. Kumar, T. S. N. Sankara Narayanan, S. Ganesh SundaraRaman and S. K. Seshadri: Tribol. Int., 2010, 43, 12451252.

    20. S. Kumar, T. S. N. Sankara Narayanan, S. Ganesh Sundara

    Raman and S. K. Seshadri: Mater. Sci. Eng. C, 2010, C30, 921

    927.

    21. A. Nevilee and B. A. B. McDougall: Proc. Inst. Mech. Eng. L:

    Mater. Des. Appl., 2002, 216, 3141.

    22. A. Berradja, F. Bratu, L. Benea, G. Willems and J. P. Celis:Wear,

    2006, 261, 987993.

    23. M. Azzi and J. A. Szpunar:Biomol. Eng., 2007, 24, 443446.

    24. J. Komotori, N. Hisamori and Y. Ohomori:Wear, 2007, 263, 412

    418.

    25. P. Ponthiaux, F. Wagner, D. Drees and J. P. Celis: Wear, 2004,

    256, 459468.

    26. F. Contu, B. Elsener and H. Bohni: Electrochim. Acta, 2004, 50,

    3341.

    27. B. W. Callen, B. F. Lowenberg, S. Lugowski, R. N. S. Sodhi and

    J. E. Davies: J. Biomed. Mater. Res., 1995, 29, 279290.28. M. Masmoudi, M. Assoul, M. Wery, R. Abdelhedi, F. el Halouani

    and G. Monteil: J. Alloys Compd, 2009, 478, 726730.

    29. B. F. Lowenberg, S. Lugowski, M. Chipman and J. E. Davies:

    J. Mater. Res. Mater. Med., 1994, 5, 467472.

    30. B. W. Callen, R. N. S. Sodhi and K. Griffiths: Prog. Surf. Sci.,

    1995, 50, 269279.

    31. E. Martin, M. Azzi, G. A. Salishchev and J. Szpunar: Tribol. Int.,

    2010, 43, 918924.

    32. N. M. Everitt, J. Ding, G. Bandak, P. H. Shipway, S. B. Leen and

    E. J. Williams: Wear, 2009, 267, 283291.

    33. M. Navarro, A. Michiardi, O. Castano and J. A. Planell:J. R. Soc.

    Interface, 2008, 5 , 11371158.

    34. C. Sedarat, M. F. Harmand, A. Naji and H. Nowzari: J.

    Periodontal Res., 2001, 36, 269274.

    35. D. G. Barceloux:J. Toxicol. Clin. Toxicol., 1999, 37, 265278.

    36. J. Black: Biological performance of materials fundamentals ofbiocompatibility, 2nd edn; 1992, New York, Marcel Dekker Inc.

    37. M. Jasty:J. Appl. Biomater., 1993, 4 , 273276.

    Sivakumar et al. Corrosion behaviour of Ti6Al4V and CP-Ti in Ringers solution

    16 4 Tribology 2011 VOL 5 NO 4