10
Electrochimica Acta 137 (2014) 280–289 Contents lists available at ScienceDirect Electrochimica Acta j our na l ho me pa g e: www.elsevier.com/locate/electacta Metastable beta Ti-Nb-Mo alloys with improved corrosion resistance in saline solution R. Chelariu a , G. Bolat b , J. Izquierdo c , D. Mareci b , D.M. Gordin d , T. Gloriant d , R.M. Souto c,e,a The “Gheorghe Asachi” Technical University of Iasi, Faculty of Materials Science and Engineering, 41 Prof. dr. doc. D. Mangeron St., 700050, Iasi, Romania b The “Gheorghe Asachi” Technical University of Iasi, Faculty of Chemical Engineering and Environmental Protection, 73 Prof. dr. doc. D. Mangeron St., 700050, Iasi, Romania c Department of Chemistry, University of La Laguna, E-38205 La Laguna, Tenerife, Canary Islands, Spain d INSA Rennes, UMR CNRS 6226 Institut des Sciences Chimiques de Rennes/Chimie Métallurgie, 20 avenue des Buttes de Coësmes, 35043 Rennes Cedex, France e University of La Laguna, Instituto Universitario de Materiales y Nanotecnologías, 38200 La Laguna (Tenerife, Canary Islands), Spain a r t i c l e i n f o Article history: Received 29 January 2014 Received in revised form 3 June 2014 Accepted 4 June 2014 Available online 10 June 2014 Keywords: Ti-Nb-Mo alloys Corrosion resistance Ti based surface oxides Saline solution Biomaterial application a b s t r a c t The present study explores the microstructural characteristics and electrochemical responses of four metastable beta Ti-Nb-Mo alloys for biomedical implantation. They were synthesized by the cold cru- cible levitation melting technique, and compositions were selected to keep the molybdenum equivalency close to 12 wt% Mo eq . For the sake of comparison, Ti12Mo was also investigated. Microstructural char- acterization reveals that all the alloys are (body-centred cubic structure), and the surface is composed by equiaxial grains with dimensions in the range of tens to hundreds m. The corrosion resistance (potentiodynamic polarization and electrochemical impedance spectroscopy) of the alloys was deter- mined in 0.9 wt% NaCl saline solution at 25 C. The materials spontaneously form a passivating oxide film on their surface, and they are stable for polarizations up to +1.0 V SCE . No evidence of localized breakdown of the oxide layers is found for polarizations more positive than those encountered in the human body. The passive layers show dielectric characteristics, and the wide frequency ranges displaying capacitive characteristics occur for both higher niobium contents in the alloy and longer exposures to the saline solution. The insulating characteristics of the oxide-covered surfaces were investigated by scanning elec- trochemical microscopy operated in the feedback mode, using ferrocene-methanol as redox mediator. Both z-approach curves and amperometric images were taken over the surface of the samples both at their open circuit potential and polarized. It has been found that Ti8Nb10Mo and Ti16Nb8Mo exhibit the lowest activity towards electron transfer. The new Ti-Nb-Mo ternary alloys are regarded to be potential candidates for biomedical application on the basis of both their microstructural characteristics and their corrosion resistance in saline solution with chloride content equivalent to body fluids. © 2014 Elsevier Ltd. All rights reserved. 1. Introduction The corrosion of implantable metallic materials remains an important issue concerning its biocompatibility and long-term sta- bility of implants. The in vivo degradation of implants leads to the loss of its functionality and integrity, and corrosion prod- ucts induces local and/or systemic effects [1–3]. The development and characterization of new Ti-based alloys containing non-toxic elements has become the topic of investigation [4–6] due to Corresponding author. E-mail addresses: [email protected], [email protected] (R.M. Souto). awareness of systemic effects of corrosion from the materials com- monly employed in the manufacture of biomedical implants for bone and teeth replacement [7,8], including reports of metallosis and necrosis [9–11]. In the search for new materials, -Ti type alloys are regarded to be the most promising ones. Thus, new -Ti alloys using Nb, Ta, Zr and Mo as alloying elements (-stabilizer elements) have been developed [12–21] owing to their low elastic modulus, high strength and corrosion resistance [22]. As alloying element, niobium is well known to be very good to resist against corrosion in simulated body fluid and presents an excellent biocompatibility [23,24]. The use of molybdenum is still controversial [25,26] but some studies have demonstrated adequate mechanical compatibility and good cyto-compatibility http://dx.doi.org/10.1016/j.electacta.2014.06.021 0013-4686/© 2014 Elsevier Ltd. All rights reserved.

Metastable beta Ti-Nb-Mo alloys with improved corrosion resistance in saline solution

  • Upload
    ull

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Mr

RTa

b

7c

d

3e

a

ARRAA

KTCTSB

1

ibtuae

h0

Electrochimica Acta 137 (2014) 280–289

Contents lists available at ScienceDirect

Electrochimica Acta

j our na l ho me pa g e: www.elsev ier .com/ locate /e lec tac ta

etastable beta Ti-Nb-Mo alloys with improved corrosionesistance in saline solution

. Chelariua, G. Bolatb, J. Izquierdoc, D. Marecib, D.M. Gordind,. Gloriantd, R.M. Soutoc,e,∗

The “Gheorghe Asachi” Technical University of Iasi, Faculty of Materials Science and Engineering, 41 Prof. dr. doc. D. Mangeron St., 700050, Iasi, RomaniaThe “Gheorghe Asachi” Technical University of Iasi, Faculty of Chemical Engineering and Environmental Protection, 73 Prof. dr. doc. D. Mangeron St.,00050, Iasi, RomaniaDepartment of Chemistry, University of La Laguna, E-38205 La Laguna, Tenerife, Canary Islands, SpainINSA Rennes, UMR CNRS 6226 Institut des Sciences Chimiques de Rennes/Chimie Métallurgie, 20 avenue des Buttes de Coësmes,5043 Rennes Cedex, FranceUniversity of La Laguna, Instituto Universitario de Materiales y Nanotecnologías, 38200 La Laguna (Tenerife, Canary Islands), Spain

r t i c l e i n f o

rticle history:eceived 29 January 2014eceived in revised form 3 June 2014ccepted 4 June 2014vailable online 10 June 2014

eywords:i-Nb-Mo alloysorrosion resistancei based surface oxidesaline solutioniomaterial application

a b s t r a c t

The present study explores the microstructural characteristics and electrochemical responses of fourmetastable beta Ti-Nb-Mo alloys for biomedical implantation. They were synthesized by the cold cru-cible levitation melting technique, and compositions were selected to keep the molybdenum equivalencyclose to 12 wt% Moeq. For the sake of comparison, Ti12Mo was also investigated. Microstructural char-acterization reveals that all the alloys are � (body-centred cubic structure), and the surface is composedby � equiaxial grains with dimensions in the range of tens to hundreds �m. The corrosion resistance(potentiodynamic polarization and electrochemical impedance spectroscopy) of the alloys was deter-mined in 0.9 wt% NaCl saline solution at 25 ◦C. The materials spontaneously form a passivating oxide filmon their surface, and they are stable for polarizations up to +1.0 VSCE. No evidence of localized breakdownof the oxide layers is found for polarizations more positive than those encountered in the human body.The passive layers show dielectric characteristics, and the wide frequency ranges displaying capacitivecharacteristics occur for both higher niobium contents in the alloy and longer exposures to the salinesolution. The insulating characteristics of the oxide-covered surfaces were investigated by scanning elec-trochemical microscopy operated in the feedback mode, using ferrocene-methanol as redox mediator.

Both z-approach curves and amperometric images were taken over the surface of the samples both attheir open circuit potential and polarized. It has been found that Ti8Nb10Mo and Ti16Nb8Mo exhibit thelowest activity towards electron transfer. The new Ti-Nb-Mo ternary alloys are regarded to be potentialcandidates for biomedical application on the basis of both their microstructural characteristics and theircorrosion resistance in saline solution with chloride content equivalent to body fluids.

© 2014 Elsevier Ltd. All rights reserved.

. Introduction

The corrosion of implantable metallic materials remains anmportant issue concerning its biocompatibility and long-term sta-ility of implants. The in vivo degradation of implants leads tohe loss of its functionality and integrity, and corrosion prod-

cts induces local and/or systemic effects [1–3]. The developmentnd characterization of new Ti-based alloys containing non-toxiclements has become the topic of investigation [4–6] due to

∗ Corresponding author.E-mail addresses: [email protected], [email protected] (R.M. Souto).

ttp://dx.doi.org/10.1016/j.electacta.2014.06.021013-4686/© 2014 Elsevier Ltd. All rights reserved.

awareness of systemic effects of corrosion from the materials com-monly employed in the manufacture of biomedical implants forbone and teeth replacement [7,8], including reports of metallosisand necrosis [9–11]. In the search for new materials, �-Ti type alloysare regarded to be the most promising ones. Thus, new �-Ti alloysusing Nb, Ta, Zr and Mo as alloying elements (�-stabilizer elements)have been developed [12–21] owing to their low elastic modulus,high strength and corrosion resistance [22].

As alloying element, niobium is well known to be very good

to resist against corrosion in simulated body fluid and presentsan excellent biocompatibility [23,24]. The use of molybdenum isstill controversial [25,26] but some studies have demonstratedadequate mechanical compatibility and good cyto-compatibility

imica

oTmc

aohdp[twmTipT�ipcwttptomibismi

atwmtn(trclw1oThwtd0t

2

2

iT

R. Chelariu et al. / Electroch

f Ti-alloys containing molybdenum such as Ti–Mo, Ti–Mo–Ta ori–Mo–Zr–Fe [27–30]. The advantage of using Mo is that this ele-ent has strong �-stabilizing properties on titanium alloys in

omparison with Nb, Ta or Zr elements [31,32].In recent years, Ti-Mo [32–35], Ti-Nb [36–39], and Ti-Nb-Mo

lloys [40–42] were investigated as candidate biomaterials basedn their good mechanical properties, shape memory behavior andigh corrosion resistance. The structure of Ti-Mo alloys stronglyepends on Mo content, � phase being stabilized at room tem-erature for Mo contents above 10 wt.% [32,33]. Yet, Zhou et al.33] identified a small volume fraction of � phase in solution-reated Ti10Mo from the XRD pattern of the alloys. Ti-Nb alloys,ith Nb contents in the interval between 3 and 35 at.% (approxi-ately 5-50 wt.%) [34–37,43] were investigated as an alternative to

i-Ni shape memory alloys [36], because nickel can easily corrodento Ni+ and Ni2+ ions of high toxicity level, allergenic nature, andossible carcinogenic effect [44–47]. The structure of the binaryi-Nb alloys varies with Nb content too, namely the hexagonal’ phase is dominant up to 15 wt.% Nb, orthorhombic �” phase

s the principal constituent between 15-30 wt.% Nb, and bcc �hase is almost entirely retained at room temperature for Nbontent above 30 wt.% [36,37,39,43,48]. The presence of � phaseas also reported for some Ti-Nb alloys with Nb content greater

han 27 wt.%. Reports on the effects of Mo and Nb contents onhe transformation temperatures, microstructures and mechanicalroperties of Ti-Nb-Mo alloys are available in the scientific litera-ure [36,40,41,43], demonstrating the strong effect of Mo contentn the stability of �-Ti phase. On the other hand, surface treat-ents can induce microstructural changes in the alloys originating

mproved mechanical properties, as reported for Ti12Mo13Nby Gabriel et al. [47], whereas de Almeida and coworkers [40]

nvestigated the effect of the aging heat treatment on the corro-ion resistance in the case of Ti12Mo13Nb and Ti12Mo20Nb. Theicrostructure of these two alloys contained both � and � phases

n their aged state [40].The aim of this study was to elaborate and to characterize new

lloy compositions in the ternary Ti-Mo-Nb system with the objec-ive to design new biomedical metastable beta Ti-based alloysith improved corrosion resistance. The effects of �-stabilizer ele-ents on the martensite transus temperature [50] and stability of

he �-phase [40] is often expressed using the so-called “molybde-um equivalent” [Moeq] [49,50], where the new alloying elementsin wt.%), such as Nb in this system, are expressed in terms ofhe equivalent weight ratio of molybdenum on the basis of theirelative atomic weights. In this work, four new Ti-Mo-Nb alloyompositions were produced in which the molybdenum equiva-ency was fixed at around 12 wt.% Moeq. That is, Ti-Nb-Mo alloys

ere designed in conditions of the same value of Moeq (about2%), and different Nb/Mo ratios obtained through the replacementf 2 wt.% Mo (2% Moeq) by 8 wt.% Nb (approximately 2% Moeq).hus, new chemical compositions were generated, but the alloysad the same value of Moeq. For the sake of comparison, Ti-12Moas also synthesized in the same manner. The microstructures of

he alloys were characterized using optical microscopy and X-rayiffraction, whereas their corrosion behaviour was investigated in.9 wt.% NaCl solution, equivalent to the saline average content inhe human body fluids [1,51].

. Material and methods

.1. Alloy synthesis

In this study, the five following alloy compositions (givenn wt.%) were synthesized: Ti12Mo, Ti10Mo8Nb, Ti8Mo16Nb,i6Mo24Nb, and Ti4Mo32Nb. They were synthesized by the cold

Acta 137 (2014) 280–289 281

crucible levitation melting (CCLM) technique in an induction fur-nace (Fives Celes, Lautenbach, France) under a pure Ar atmosphere,which was introduced after several cycles of high vacuum pumping.After solidification, the alloys were heat treated at 950 ◦C for 20hours (annealed in the �-phase domain) in order to homogenize themicrostructure due to the thermal gradient occurring during solid-ification, and then water quenched. The ingots were cold rolled to90% in thickness and machined to obtain tensile test specimensand corrosion test samples. After this mechanical treatment, allsamples were solution treated at 850 ◦C for 30 minutes and waterquenched. The aim of this treatment was to restore a fully recrys-tallized metastable � microstructure from the cold-rolled state.

2.2. Alloy characterization methods

To check the nominal chemical composition of each alloy afterfabrication, EDX analyses were carried out in a JEOL JSM 6400 scan-ning electron microscope operating at 20 kV.

The crystalline structure of the recrystallized alloys was charac-terized by X-ray diffraction (XRD) at ambient temperature using aPhilips PW 1830/00 (Eindhoven, The Netherlands) diffractometer

(CuK�1 radiation, 1.5406 ´A wavelength).The microstructures of the alloys were characterized by optical

microscopy using a Leica DM RM system (Wetzlar, Germany). Priorto optical imaging, the samples were first mechanically abradedusing a sequence of silicon carbide abrasive papers followed bya final polishing step with a colloidal silica suspension (particlesize: 50 nm). Next, the samples were cleaned in alcohol and dis-tilled water using an ultrasonic bath, and finally etched in a solutioncontaining 5 wt.% HNO3, 5 wt.% HF and the balance in H2O. Sampleswere observed under polarized light to enhance the phase contrast.

2.3. Electrochemical measurements

The test specimens were placed in a glass corrosion flow cell kit(C145/170, Radiometer, France), which was filled with naturally-aerated 0.9 wt.% NaCl solution [52]. A saturated calomel electrodewas used as the reference electrode, and a platinum coil as thecounter electrode. The potentials in this paper are reported versusthe saturated calomel electrode (SCE). The temperature of theelectrochemical cell was maintained at 25 ± 0.1 ◦C. Though theintra-oral temperature fluctuates on ingestion of hot or cold foodand beverage, it can be reasonably approximated in experimentsby setting the environmental temperature at 25 ◦C [53].

Electrochemical measurements were performed using a poten-tiostat model PARSTAT 4000 (Princeton Applied Research, NJ, USA).The instrument was controlled by a personal computer and Ver-saStudio software. Prior to testing, the working electrodes weremechanically abraded using emery paper up to 2000 grit, next pol-ished with 0.3 �m alumina suspension, ultrasonically cleaned inacetone and deionized water, and finally dried in open air.

Potentiodynamic polarization and electrochemical impedancespectroscopy (EIS) methods were employed to characterize thecorrosion behaviour of these new titanium-based alloys. The poten-tiodynamic polarization tests were initiated after 1 h immersion ofthe samples in the saline solution, when the open circuit poten-tial acquired a stable value. The tests were conducted by scanningthe potential, at the rate of 0.5 mV s−1, from -1.0 VSCE up to +1.0VSCE, followed by potential reversal down to +0.5 VSCE, in orderto explore the eventual occurrence of pitting. Standard techniqueswere used to extract the zero current potential (Ecorr), and the

corrosion current density (jcorr), namely the application of Tafelanalysis for a range of ±100 mV around the open circuit poten-tial. From the measured potentiodynamic polarization curves, thepassive current density (jpass) was determined too.

2 imica Acta 137 (2014) 280–289

qacaottZ

rd

2

(aaTatstfrescbidwmtmtstnzmdttbttt2n(

3

3

cot

tit

Fig. 1. XRD patterns of Ti12Mo and Ti-Mo-Nb alloys possessing a common molyb-denum equivalency composition of ca. 12 wt.% Moeq.

82 R. Chelariu et al. / Electroch

Electrochemical impedance spectra were measured over a fre-uency range extending from 100 kHz to 10 mHz using a 10 mVmplitude AC voltage signal. The EIS tests were recorded at the openircuit potential developed by the samples after 10 minutes, 1 hour,nd 24 hours of immersion in the saline solution. EIS tests werebtained to evaluate the characteristics of passive films formed onhe surface of the alloys under investigation. Analysis of the spec-ra was performed in terms of equivalent circuit (EC) fitting usingSimpWin software.

All electrochemical tests were repeated three times to ensureeproducibility of the measurements, and reproducibility of theata is expressed in terms of the corresponding standard deviation.

.4. Scanning electrochemical microscopy (SECM)

A high-resolution SECM equipment supplied by SensolyticsBochum, Germany), was employed for spatially-resolved char-cterization of the electrochemical reactivity of the TiNbMolloys. The instrument was built around a PalmSens (Utrecht,he Netherlands) electrochemical interface, all controlled with

personal computer. The alloy specimens were either tested atheir open circuit potentials spontaneously developed in the salineolution, or polarized using the bipotentiostat built in the elec-rochemical interface. The specimens were mounted horizontallyacing upwards at the bottom of a cell made of polytetrafluo-oethene, which was equipped with an Ag/AgCl/(3 M) KCl referencelectrode and a platinum counter electrode. For the sake of con-istency, potentials were subsequently referred to the standardalomel electrode by taking in account the potential differenceetween the two electrodes. The electrochemical cell was located

nside a Faraday cage. Tip microelectrodes were made from 10 �mia. platinum wires sealed in glass. 0.5 mM ferrocene-methanolas added to the 0.9 wt.% NaCl solution to act as electrochemicalediator at the tip. To enable the oxidation of ferrocene-methanol

he tip was kept at a constant potential of +0.46 VSCE [54]. Theicromanipulator stand of the SECM instrument was used to hold

he microelectrode in place. The operating tip distance over theample for SECM imaging was established by slowly approachinghe surface of the alloy specimen with the tip and simulta-eously recording the measured current at the microelectrode vs.

displacement (i.e., z-approach curve). In the experiments theicroelectrode was either stopped when the measured current

ecreased to 30% of the steady-state value in the bulk of the elec-rolyte, or when a current peak was detected while approachinghe surface, and indication of mixed positive and negative feed-ack behaviours. Subsequently, the tip was withdrawn 10 �m fromhe sample surface, and images were obtained by scanning theip parallel to the sample surface (i.e., constant height opera-ion). SECM images were recorded rastering an area of 250 �m x50 �m using a scan rate of 30 �m s−1. Temperature control wasot performed, and data were recorded at ambient temperatureca. 22 ± 3 ◦C)

. Results and discussion

.1. Microstructural characterization

The chemical compositions of the etched Ti-based alloys wereonfirmed by EDS analysis, and each nominal composition wasbtained with less than 0.5 wt.% allowance, which corresponds tohe precision of the EDS method.

The five alloys were found to be � (body-centred cubic struc-ure) by X-ray diffraction as shown in Fig. 1. Indeed, all the peaksn the diffractograms were unambiguously indexed according tohis crystallographic structure, and the corresponding diffracted

Fig. 2. Optical micrograph of the Ti10Mo8Nb alloy showing the recrystallized �-grain microstructure.

planes are indicated in the plot. The occurrence of some small addi-tional peaks was observed for the alloys rich in Mo (i.e., Ti12Mo,Ti10Mo8Nb, and Ti8Mo16Nb). In fact, the presence of hexago-nal �ath phase nanoprecipitates (dimension less than 5 nm) inmetastable � Ti-Mo based alloys, which are formed by displacedtransformation during quenching, has been reported for a long time[55–57].

The micrograph depicted in Fig. 2 shows the microstructure ofthe Ti10Mo8Nb alloy. � equiaxial grains with dimensions in therange of tens to hundreds �m are observed, in good agreement withthe XRD observations. Very similar microstructures were observedfor the other alloys, and are not shown here.

3.2. Electrochemical characterization: corrosion resistance

Fig. 3 depicts the potentiodynamic polarization curves mea-sured for the various alloys during immersion in naturally-aeratedsaline solution at 25 ◦C. All samples depicted similar features, char-acteristic of materials exhibiting a passive behaviour in the testenvironment, though there is no evidence of an active-passive tran-sition [58]. This feature indicates that the oxide film spontaneouslydeveloped at the surface of the alloys upon immersion in the testelectrolyte exhibits passivation characteristics, a feature previously

observed for titanium metal [59] and Ti-based alloys such as nitinol[60], Ti-Nb-Hf [61] and ZrTi alloys [62]. Additionally, the poten-tial excursion up to +1.0 VSCE, a value significantly higher thanthe polarizations ever measured in the human body [63], does not

R. Chelariu et al. / Electrochimica

Fig. 3. Potentiodynamic polarization curves for the Ti12Mo and Ti-Mo-Nb alloysafter 1 h immersion in naturally aerated 0.9 wt.% NaCl solution at 25 ◦C. Scanningrate: 0.5 mV s−1.

Table 1Corrosion parameters (and standard deviation values) determined from the poten-tiodynamic polarization curves measured for the Ti-Nb-Mo alloys in naturallyaerated 0.9 wt.% NaCl solution.

Alloy Ecorr, VSCE jcorr, �A cm−2 jpass, �A cm−2

Ti12Mo -0.701 (0.021) 0.56 (0.07) 2.24 (0.04)Ti8Nb10Mo -0.640 (0.022) 0.43 (0.05) 2.32 (0.05)

soptcwTmVbtgvttmtt

Ti16Nb8Mo -0.609 (0.021) 0.41 (0.04) 1.24 (0.04)Ti24Nb6Mo -0.514 (0.020) 0.38 (0.04) 0.86 (0.03)Ti32Nb4Mo -0.451 (0.019) 0.24 (0.05) 0.66 (0.04)

how the breakdown of the passive films related to the initiationf localized corrosion processes in any of the alloys. The corrosionotential (Ecorr) and the corrosion current density (jcorr) values ofhe materials were determined by Tafel analysis of both anodic andathodic branches of the polarization curves using CorrView soft-are. The mean values and standard deviations are compiled in

able 1, together with the passivation current densities (jpass) deter-ined from the current plateaus at potentials more positive than 0

SCE. Since the values of the corrosion current densities determinedy the Tafel method are smaller than the passivation current densi-ies for all the alloys, the material is passive at Ecorr, and the valuesiven in Table 1 correspond to a mixed potential. Accordingly, thealues of the corrosion current density (jcorr) do not correspondo the actual corrosion rate of the alloys due to the presence of

he passive films on their surface, thus the values given in Table 1

ay be used to provide a relative ranking of the corrosion resis-ance. From these data it is possible to observe that the values ofhese electrochemical parameters support that the decrease in the

Fig. 4. Nyquist and Bode diagrams for Ti12Mo and Ti-Mo-Nb alloys after 10 m

Acta 137 (2014) 280–289 283

molybdenum content, and the subsequent increase in niobium con-tent though made in such a way that the equivalent molybdenumcontent could be maintained constant, effectively leads to the mea-surement of nobler values for Ecorr, accompanied by smaller valuesfor the current densities jcorr and jpass. This observation revealsan improvement in the corrosion protection characteristics of thealloys by increasing the content of niobium [64]. The same obser-vation was found by Atapour et al. [65] in their comparison ofthe corrosion behaviours of Ti6Al4V, Ti13Mo7Zr3Fe (both (� + �)and metastable � condition), and Ti35Nb7Zr5Ta alloys performedin 0.9 wt.% NaCl solution at 37 ◦C. They reported smaller jcorr val-ues for Ti13Mo7Zr3Fe (metastable � condition) and Ti35Nb7Zr5Taas compared to those for Ti6Al4V and Ti13Mo7Zr3Fe ((� + �) con-dition), whereas Ti35Nb7Zr5Ta showed the smallest passivationcurrent density (jpass). On the other hand, the current density val-ues determined by Zheng et al. [66] for Ti–16 at.% Nb-(4.0, 4.5, 5.0)at.% Sn (approximately Ti-25 wt.% Nb-(8, 9, 10) wt.% Sn) are veryclose to those found in this work (namely 0.1 ≤ jcorr ≤ 0.3 �A cm−2,and 4.3 ≤ jpass ≤ 10.0 �A cm−2), though the corresponding valuesof Ecorr were slightly more positive (in the range -0.2 > Ecorr > -0.4 VSCE). A similar trend in the values of the corrosion parameterswas found for Ti-22 at.% Nb-(0, 2, 4, 6) at.% Hf (approximatelyTi- (35, 34, 32, 31) wt.% Nb-(0, 6, 11, 16) wt.% Hf) in 0.9 wt.%NaCl neutral solution maintained at 37 ◦C [67]. In this case, thecorresponding data were: 0.633 ≥ jcorr ≥ 0.052 �A cm−2, 8.0 ≥jpass ≥ 5.5 cm−2, and -0.2 > Ecorr > -0.4 VSCE. No breakdown potentialwas identified in the potentiodynamic polarization curves for thetested potential range, and no pits were observed for the retrievedsamples.

In order to obtain further information on the characteristicsof the passive films and the nature of the electrochemical pro-cesses at the interface, EIS tests were performed. The EIS data ofthe alloys measured at three different immersion times in the testsolution are presented in Figs. 4-6 in the form of Nyquist (complexversus real components of the impedance) Bode plots (impedancemodulus and phase angle versus frequency diagrams). In all cases,only one time constant is observed in the spectra, and they couldbe satisfactorily simulated using the simplified Randles’ circuitshown in Fig. 7A, which assumes that the corrosion of the passivemetal is hindered by an oxide film that acts as a barrier-type com-pact layer. This equivalent circuit has been previously employedto describe the impedance behaviour of various titanium-basedalloys exposed to saline and physiological solutions [3,67–74]. Yet,most oxide layers formed on passivating metals, including mosttitanium-based alloys [75,76], usually present a duplex structure,with a thin compact inner layer and a somewhat more porous exter-

nal layer that may be eventually penetrated by the electrolyte [77].The latter accounts for the observation of a second time constantin the impedance spectra, and the impedance spectra are mod-elled by the equivalent circuit given in Fig. 7B corresponding to

in in naturally aerated 0.9 wt.% NaCl solution at 25 ◦C, measured at Ecorr.

284 R. Chelariu et al. / Electrochimica Acta 137 (2014) 280–289

Fig. 5. Nyquist and Bode diagrams for Ti12Mo and Ti-Mo-Nb alloys after 1 h in naturally aerated 0.9 wt.% NaCl solution at 25 ◦C, measured at Ecorr.

er 24 h

atpftfobsotfitafvaw

Fa[

from fitting of the Rsol(RpLQpL) equivalent circuit to the EIS data aregiven in Table 2. The fitting quality of the experimental EIS spectrato the model was established from the chi-squared (�2) values,

Fig. 6. Nyquist and Bode diagrams for Ti12Mo and Ti-Mo-Nb alloys aft

n unsealed inhomogeneous surface film [78]. This correspondso the barrier characteristics of the oxide film towards electrolyteenetration that dominates the impedance response in the low-requency range (described by the time constant Rox Qox), and tohe charge transfer process in the surface of the alloys at higherrequencies (represented by Rct and Qdl). Indeed, the formationf oxide layers on TiMo alloys presenting a duplex structure haseen previously reported by our group [79]. This apparent incon-istency can be explained by considering that passive films formedn the alloys are rather thin (typically in the order of a few nanome-ers) [79], and in the event that the pores in the external layer arelled with electrolyte, the contribution from this porous layer tohe overall electrochemical behaviour of the system is very small,nd the impedance response is dominated by the inner layer. In

act, the values of RpL found in this work are close to those pre-iously reported for the inner compact layer in TiMo alloys [79],nd certainly too small for a perfect compact barrier layer. In thisay, the passive film is considered to present a duplex structure,

ig. 7. Equivalent circuits (EC) used to analyze the impedance spectra: (A) model of compact barrier film; and (B) two-layer model of an unsealed porous surface film78].

in naturally aerated 0.9 wt.% NaCl solution at 25 ◦C, measured at Ecorr.

though corrosion protection almost exclusively arises from the bar-rier properties exhibited by the inner oxide layer.

The parameters RpL and QpL describe the properties of thepassive films formed on these metallic materials, respectively theresistance and capacitance of the compact oxide layers, whereasRsol stands for the solution resistance of the test electrolyte. Theuse of a constant phase element (CPE, Q) was required to accountfor the distributed relaxation feature commonly exhibited by thepassive films formed on titanium-based alloys [80]. It arises fromsurface inhomogeneities (roughness and defects) present at themicroscopic level [1,68]. The impedance parameters determined

Table 2Parameters of the equivalent circuit for the Ti-Nb-Mo alloys in naturally aerated0.9 wt.% NaCl solution for various durations of immersion. AC polarization wasapplied around their corresponding open circuit potential values in the electrolytes.

Alloy Rsol, � cm2 QpL, �S cm−2 sn n RpL, M� cm2

10 minutes immersion timeTi12Mo 59.6 0.13 0.82 0.58Ti8Nb10Mo 52.3 0.13 0.83 0.62Ti16Nb8Mo 66.1 0.12 0.83 0.75Ti24Nb6Mo 51.3 0.11 0.84 0.97Ti32Nb4Mo 61.9 0.11 0.85 1.101 hour immersion timeTi12Mo 59.6 0.12 0.83 0.82Ti8Nb10Mo 52.3 0.12 0.83 0.95Ti16Nb8Mo 65.1 0.12 0.83 0.98Ti24Nb6Mo 54.3 0.11 0.84 1.20Ti32Nb4Mo 61.2 0.08 0.86 1.6024 hours immersion timeTi12Mo 60.4 0.11 0.84 0.94Ti8Nb10Mo 52.0 0.11 0.84 1.10Ti16Nb8Mo 66.3 0.11 0.84 1.10Ti24Nb6Mo 61.1 0.08 0.86 1.30Ti32Nb4Mo 62.2 0.07 0.86 1.90

imica Acta 137 (2014) 280–289 285

ie

irttsblau-aetmttdiaittaN

3r

adccptaittiarfse2wtrt

ifbvtmdstss

Fig. 8. Typical z-approach normalized curves measured above Ti12Mo and Ti-Nb-

R. Chelariu et al. / Electroch

n the order of 10−4, indicating the good agreement betweenxperimental and simulated data.

From the inspection of the Bode plots, it is observed that thempedance characteristics of the system in the high frequencyange are exclusively determined by the resistance of the elec-rolyte. That is, the phase angle drops to zero degrees whereashe impedance modulus tends to ca. 60 � cm2. This value of theolution resistance (Rsol) is almost half of that previously reportedy Wang et al. in the same electrolyte [67]. In the medium and

ow frequency ranges, the Bode-phase plots exhibit a plateau withngles close to -80o, whereas the values of the impedance mod-lus varied linearly with the frequency presenting slopes close to0.9. These features are characteristic of passive films presenting

highly capacitive behaviour [67,81,82]. This capacitive behaviourxtends over a wider frequency range as the alloys are exposed tohe test electrolyte for longer exposures, confirming the improve-

ent of the barrier characteristics of the passive film formed onhe Ti-Nb-Mo alloys in this electrolytic environment. Additionally,he reported capacitive behaviour of passive films is more evi-ent as the content of Nb increases and content of Mo decreases

n the materials. In summary, the data presented in Table 2 shown enhancement of the corrosion resistance of the alloys with thencrease of both the Nb content in the alloy and the immersion timeo the saline solution. And these values of the impedance parame-ers RpL, n, and QpL are close to those previously reported for Ti-22t.% Nb-(0, 2, 4, 6) at.% Hf (approximately Ti-(35, 34, 32, 31) wt.%b-(0, 6, 11, 16) wt.% Hf) in 0.9 wt.% NaCl neutral solution [67].

.3. Scanning microelectrochemical characterization: surfaceeactivity of the passive films

SECM measurements were conducted as both vertical z-pproach curves and 2D grid scans. In both cases, various randomlyistributed portions of the exposed surface of the alloys wereonsidered, in order to get a fair representation of the electro-hemical behaviour of the material at the micrometer scale. Tipotential was set at +0.46 VSCE to ensure diffusional control forhe reversible oxidation of ferrocene-methanol, a chemical thatcted as the redox mediator. The experimental sequence consistedn firstly measuring the z-approach curves at four different loca-ions of the surface, by bringing the tip from the bulk electrolyteowards the surface while measuring the faradaic current flow-ng at the tip for ferrocene-methanol oxidation. This procedurelso allowed correcting any eventual tilt of the substrate occur-ing during assembling of the small microelectrochemical cell usedor the SECM measurements. Next, the tip was retracted to a tip-ubstrate distance equal to the tip diameter (a = 10 �m) to measurelectrochemical activity maps at constant height over a 250 �m x50 �m area. Surface activity towards electron transfer reactionsas evidenced by higher tip current values while scanning than in

he bulk of the electrolyte, resulting from the capacity of the surfaceegion to donate one electron to the ferrocinium ion generated athe tip.

Fig. 8 shows typical z-approach curves measured shortly aftermmersion of the five alloys in the test solution containingerrocene-methanol. Normalized current and distance values haveeen employed for easy comparison of the results, so every currentalue is referred to the limiting current measured in bulk solu-ion, and distances are referred to the diameter of the platinum

icrodisc. Alloy surfaces were kept unpolarized in the test solutionuring these measurements. It has been observed that the Ti12Mo

urface presents the higher electron transfer rate, as evidenced byhe big increase of the tip current when it approaches to the sub-trate (i.e., positive feedback behaviour). Thus, the passive layerpontaneously formed on the surface of TiMo under spontaneous

Mo alloys during immersion in 0.5 mM ferrocene-methanol + 0.9 wt.% NaCl solutionat ambient temperature. The samples were left unbiased at their corresponding Ecorr.Tip potential: +0.46 VSCE.

open circuit potential conditions, maintains a rather high electricalconductivity and can operate as a source of electrons for the regen-eration of ferrocene-methanol species from the ferrocinium ionsgenerated at the tip. It must be realized that the passive film con-stitutes a barrier to ion transfer but not to electron transfer, andredox electron transfer reactions can occur on the oxide-coveredsurface and not exclusively when the metal is free from any surfacefilm [54].

Redox mediator regeneration becomes more difficult when thetip approaches the surfaces of Ti8Nb10Mo and Ti16Nb8Mo instead,as evidenced by the onset of a mixed kinetic behaviour. That is,though the tip current initially increases while the tip approachesthe surface from the bulk electrolyte, subsequent blockage to dif-fusion of the redox mediator towards the tip by a somewhatinsulating surface at small tip-substrate distances leads to a finaldecrease of the tip current. It is shown that the passive layer is notan efficient electron donor for the reduction of ferrocinium. Finally,the two alloys with the highest niobium contents (i.e., Ti24Nb6Moand Ti32Nb4Mo) show positive feedback behaviour, thus closerresembling the electrochemical behaviour found for Ti12Mo. Theseobservations on the electrochemical activity of the alloys at themicrometer scale apparently contradict the average measurementsbased on EIS measurements. Impedance data showed a systematicgrowth of the corrosion resistance of the oxide films for increas-ing contents of niobium in the alloys (cf. Table 2), and this mightbe regarded to directly correlate to the susceptibility of the sur-faces towards electron exchange. However, according to the SECMresults in Fig. 8, it seems that there is an optimum niobium con-tent for the alloys, around 8 to 16 wt.%, if electron donation is to beavoided. Materials with the higher niobium content are effectivelymore resistant towards corrosion, but their passive layers can stillsustain electron transfer reactions at the surface. Considering alsosurface conductivity affects to the results obtained by SECM in thisconfiguration, apparently the higher protective layers formed onthe 24 and 32 wt.% niobium containing alloys are less electrically-insulating than those for 8 and 16 wt.% composition, althoughthey provide a greater protection against corrosion in the salinesolution.

The spatial distribution of electrochemical activity of the fivealloys immersed in the saline solution was evaluated by recordingSECM grid scans at constant height while the samples were bothat their open circuit potentials and polarized at -0.54, -0.24 and

+0.16 VSCE. Typical SECM images are given in Figs. 9-11 for Ti12Mo,Ti8Nb10Mo and Ti16Nb8Mo, and Ti24Nb6Mo and Ti32Nb4Mo,respectively. In general, the current measured at the tip decreasesas the potential of the substrates was set at more positive values.

286 R. Chelariu et al. / Electrochimica Acta 137 (2014) 280–289

Fig. 9. SECM images of the Ti12Mo surface while (A) left at its spontaneous OCP, and biased at (B) -0.54, (C) -0.24, and (D) +0.16 VSCE during immersion in 0.5 mM ferrocene-methanol + 0.9 wt.% NaCl solution at ambient temperature. Tip to substrate distance: 10 �m. Tip potential: +0.46 VSCE. Scan rate: 30 �m s−1.

Fig. 10. SECM images of the Ti8Nb10Mo (left) and Ti16Nb8Mo (right) surfaces while (A) left at its spontaneous OCP, and biased at (B) -0.54, (C) -0.24, and (D) +0.16 VSCE duringimmersion in 0.5 mM ferrocene-methanol + 0.9 wt.% NaCl solution at ambient temperature. Tip to substrate distance: 10 �m. Tip potential: +0.46 VSCE. Scan rate: 30 �m s−1.

R. Chelariu et al. / Electrochimica Acta 137 (2014) 280–289 287

F (A) lei eratur

TsmantAmytatctsnTtrtbtscfp

auht1tctp

ig. 11. SECM images of the Ti16Nb6Mo (left) and Ti32Nb4Mo (right) surfaces whilemmersion in 0.5 mM ferrocene-methanol + 0.9 wt.% NaCl solution at ambient temp

herefore, the oxidized ferrocinium ion generated at the tip whilecanning the alloys is no longer reduced to regenerate ferrocene-ethanol at the surface of the alloys when more anodic potentials

re applied to them. This is mainly due to the increase of the thick-ess of the passive layer when the potential increases. However,he magnitude of this effect differs from one material to another.s it can be seen in Fig. 9, Ti12Mo is able to regenerate ferrocene-ethanol at most of its surface when left at its spontaneous OCP

et clear heterogeneous behaviour is encountered. When poten-ial is set at -0.54 VSCE, normalized current values close to unityre observed. This indicates that the surface activity towards elec-ron donation has decreased, yet the positive feedback effect almostompensates the diffusion blockage of the mediator produced byhe small electrolyte volume comprised between the tip and theurface. At more positive potentials, namely -0.24 and +0.16 VSCE,egative feedback behaviour dominates the response of the system.hat is, diffusion blockage by a predominantly insulating surface ishe main behaviour, and this effect is stronger at some distributedegions where the current depletion occurs in a bigger extent. Yet,he measurements at +0.16 VSCE must be taken with some caution,ecause the potential of the substrate is already positive enoughhat some oxidation of ferrocene-methanol may occur on the sub-trate. In that event, smaller tip currents would result from theompetition between the substrate and the tip for the oxidation oferrocene-methanol, and not exclusively from negative feedbackroperties of the surface.

Smaller tip currents related to feedback are found when thelloys contain niobium measured at their corresponding OCP val-es. It must be noticed that the SECM images depicted in Fig. 10ave been plotted with a current range significantly narrower thanhose in Fig. 9, because normalized current values higher than.15 are not found. A more electrochemically-homogeneous sys-

em is encountered while scanning these surfaces, especially whenonsidering the Ti8Nb10Mo alloy. Again, the contribution of posi-ive feedback is reduced when the substrate potential is set moreositive. Yet, some differences between these two alloys may be

ft at its spontaneous OCP, and biased at (B) -0.54, (C) -0.24, and (D) +0.16 VSCE duringe. Tip to substrate distance: 10 �m. Tip potential: +0.46 VSCE. Scan rate: 30 �m s−1.

deduced from the inspection of Fig. 10C, that corresponds to thepolarization value -0.24 VSCE. Significantly smaller tip current wereobserved in the case of Ti16Nb8Mo. This may suggest that the pas-sive layer formed on this alloy during anodic polarization is moreinsulating, and this material may be regarded to be more corrosionresistant in an oxidizing environment.

Finally, images taken over the two high niobium containingalloys, presented in Fig. 11, reproduce the same behaviour than thatfound for the Ti12Mo at the OCP and -0.54 VSCE. The more protec-tive passive layer formed over Ti32Nb4Mo surface provides higherresistance towards electron transfer in some areas, and this abil-ity is maintained when biased at -0.54 VSCE, though this potentialvalue is more negative than the spontaneous OCP of the materialin the solution (cf. Table 1), but not cathodic enough to effectivelyreduce the oxide film formed on the surface of the alloy at its OCPwhile recording the z-approach curves and the SECM image given inFig. 11A. Passivated surfaces still participate in electron exchangefor ferrocinium reduction when anodic polarization increases upto -0.24 VSCE, showing a heterogeneous distribution of the elec-trochemical activity, with some areas even leading to tip currentshigher than those determined in the bulk solution. Finally, at +0.16VSCE, normalized current values are smaller for the 24 wt.% niobiumalloy, whereas a distinctive heterogeneous distribution is observedfor the material with the highest content of niobium (cf. Fig. 11D).In this case, strong positive feedback behaviour was found at thebeginning of the scan, possibly due to breakdown of the passivelayer, and the current subsequently decreased rapidly due to pas-sive layer rebuilding and the onset of competition effect towardsferrocene-methanol oxidation between the tip and the substrate.

4. Conclusions

In this study, the effect of niobium additions to Ti-Mo alloys,keeping the molybdenum equivalency close to 12 wt.% Moeq, havebeen investigated with respect to the corrosion resistance of the

2 imica

mfpic(

A

AbEobP

R

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

88 R. Chelariu et al. / Electroch

aterials in 0.9 wt.% NaCl saline solution at 25 ◦C. The observationsrom combined conventional electrochemical analysis (namelyotentiodynamic polarization measurements and electrochemical

mpedance spectroscopy (EIS)) and scanning microelectrochemicalharacterizations (using the scanning electrochemical microscopeSECM)) have led to the following conclusions:

All the investigated Ti-Nb-Mo alloys passivated spontaneouslyafter immersion in 0.9 wt.% NaCl and passive films have a capac-itive behaviour analogously to Ti12Mo.A significant improvement of the corrosion resistance of thematerials in the saline solution is achieved by the addition of nio-bium to Ti-Mo alloys. Potentiodynamic polarization data showthat both corrosion and passive current densities decrease whenthe Nb content increases.The EIS results indicate that the passive oxide film formed on thealloys is rather compact, characteristic of a sealing barrier layerexhibiting only one time constant in the spectra. The impedancevalues of the barrier layer are higher in the niobium-containingalloys than those of the binary Ti-Mo alloy with the same molyb-denum equivalent, demonstrating that the passive oxide layerensures the alloys a good corrosion resistance.The corrosion resistance provided by these oxide layers alsoimproves with increasing immersion time in the test electrolyte,at least within the first 24 hours, consistent with thickening ofthe surface films.The improved corrosion resistance can be attributed to hindranceof electron transfer on the passive film-covered metal surface. Infact, electron transfer at the surface of the passivated alloys ishindered for niobium contents is in the range of 8 to 16 wt.%, andthis behaviour is even more pronounced when positive poten-tials are applied to the substrate while immersed in the salineenvironment, as demonstrated by scanning microelectrochem-ical characterization. That is, redox reactions still occur on theoxide-covered surface of the metal though at a slower rate thanon the metal surface without any film.

cknowledgments

This work was supported by a grant of the Romanian Nationaluthority for Scientific Research, CNCS-UEFISCDI, project num-er PN-II-ID-PCE-2011-3-0218, and by the Spanish Ministry ofconomy and Competitiveness, and the European Regional Devel-pment Fund (Project No. CTQ2012-36787). A Grant awarded to J.I.y the Spanish Ministry of Education (Programa de Formación deersonal Investigador) is gratefully acknowledged.

eferences

[1] S.L. Assis, S. Wolynec, I. Costa, Corrosion characterization of titanium alloys byelectrochemical techniques, Electrochimica Acta 51 (2006) 1815–1819.

[2] A. Sargeant, T. Goswami, Hip implants: Paper V. Physiological effects, MaterialsDesign 27 (2006) 287–307.

[3] C.V. Vidal, A.I. Munoz, Electrochemical characterisation of biomedical alloysfor surgical implants in simulated body fluids, Corrosion Science 50 (2008)1954–1961.

[4] M. Long, H.J. Rack, Titanium alloys in total hip replacement, a materials scienceperspective, Biomaterials 19 (1998) 1621–1639.

[5] M. Niinomi, Mechanical properties of biomedical titanium alloys, MaterialsScience and Engineering A 243 (1998) 231–236.

[6] Y. Okazaki, S. Rao, Y. Ito, T. Tateishi, Corrosion resistance, mechanical properties,corrosion fatigue strength and cytocompatibility of new Ti alloys without Aland V, Biomaterials 19 (1998) 1197–1215.

[7] A. Balamurugan, S. Rajeswari, G. Balossier, A.H.S. Rebelo, J.M.F. Ferreira, Cor-rosion aspects of metallic implants - An overview, Materials and Corrosion 59

(2008) 855–869.

[8] O. Addison, A.J. Davenport, R.J. Newport, S. Kalra, M. Monir, J.F.W. Mosselmans,D. Props, R.A. Martin, Do ‘passive’ medical titanium surfaces deteriorate inservice in the absence of wear? Journal of the Royal Society Interface 9 (2012)3161–3164.

[

[

Acta 137 (2014) 280–289

[9] C.P. Case, V.G. Langkamer, C. Jamec, M.R. Palmer, A.J. Kemp, P.F. Heap, L.Solomon, Widespread dissemination of metal debris from implants, Journalof Bone & Joint Surgery 76B (1994) 701–712.

10] I. Milosev, V. Antolic, A. Minovic, A. Cör, S. Herman, V. Pavlovcic, P. Campbell,Extensive metallosis and necrosis in failed prostheses with cemented titanium-alloy stems and ceramic heads, Journal of Bone & Joint Surgery 82B (2000)352–357.

11] P. Korovessis, G. Petsinis, M. Repanti, T. Repantis, Metallosis after contemporarymetal on metal total hip arthroplasty. Five to nine year follow up, Journal ofBone & Joint Surgery American 88 (2006) 1183–1191.

12] D. Kuroda, M. Niinomi, M. Morinaga, Y. Kato, T. Yashiro, Design and mechanicalproperties of new beta type titanium alloys for implants materials, MaterialsScience and Engineering A 243 (1998) 244–249.

13] M. Niinomi, T. Hattori, K. Morikawa, T. Kasuga, A. Suzuki, H. Fukui, S. Niwa,Development of low rigidity �-type titanium alloy for biomedical applications,Materials Transactions 43 (2002) 2970–2977.

14] D.M. Gordin, T. Gloriant, Gh. Nemtoi, R. Chelariu, N. Aelenei, A. Guillou, D. Ansel,Synthesis, structure and electrochemical behavior of a beta Ti-12Mo-5Ta alloyas new biomaterial, Materials Letters 59 (2005) 2936–2941.

15] J. Qazi, H.J. Rack, Metastable beta titanium alloys for orthopedic applications,Advanced Engineering Materials 7 (2005) 993–998.

16] L.M. Elias, S.G. Schneider, S. Schneider, H.M. Silva, F. Malvisi, Microstructuraland mechanical characterization of biomedical Ti-Nb-Zr(-Ta) alloys, MaterialsScience and Engineering A 432 (2006) 108–112.

17] T. Gloriant, G. Texier, F. Prima, D. Laillé, D.M. Gordin, I. Thibon, D. Ansel, Syn-thesis and phase transformations of beta metastable Ti-based alloys containingbiocompatible Ta, Mo and Fe beta-stabilizer elements, Advanced EngineeringMaterials 8 (2006) 961–965.

18] N.T.C. Oliveira, S.R. Biaggio, P.A.P. Nascente, S. Piazza, C. Sunseri, F. Di Quarto,The effect of thickness on the composition of passive films on a Ti–50Zr at%alloy, Electrochimica Acta 51 (2006) 3506–3515.

19] M. Niinomi, Mechanical biocompatibilities of titanium alloys for biomedicalapplications, Journal of the Mechanical Behaviour of Biomedical Materials 1(2008) 30–42.

20] S.L. Assis, S. Wolynec, I. Costa, The electrochemical behaviour of Ti-13Nb-13Zralloy in various solutions, Materials and Corrosion 59 (2008) 739–743.

21] N.T.C. Oliveira, A.C. Guastaldi, Electrochemical stability and corrosion resis-tance of Ti–Mo alloys for biomedical applications, Acta Biomaterialia 5 (2009)399–405.

22] Y. Li, C. Yang, H. Zhao, S. Qu, X. Li, Y. Li, New developments of Ti-based alloysfor biomedical applications, Materials 7 (2014) 1709–1800.

23] Y. Okazaki, S. Rao, T. Tateishi, Y. Ito, Cytocompatibility of various metal anddevelopment of new titanium alloys for medical implants, Materials Scienceand Engineering A 243 (1998) 250–256.

24] H. Matsuno, A. Yokoyama, F. Watari, M. Uo, T. Kawasaki, Biocompatibility andosteogenesis of refractory metal implants Ti, Hf, Nb, Ta, Rh, Biomaterials 22(2001) 1253–1262.

25] E. Eisenbarth, D. Velten, M. Müller, R. Thull, J. Breme, Biocompatibility of beta-stabilizing elements of titanium alloys, Biomaterials 25 (2004) 5705–5713.

26] Y. Li, C. Wong, J. Xiong, P. Hodgson, C. Wen, Cytotoxicity of titanium and tita-nium alloying elements, Journal of Dental Research 89 (5) (2010) 493–497.

27] W.F. Ho, C.P. Ju, J.H. Chern Lin, Structure and properties of cast Ti-Mo alloys,Biomaterials 20 (1999) 2115–2122.

28] L. Trentani, F. Pelillo, F.C. Pavesi, L. Ceciliani, G. Cetta, A. Forlino, Evaluation of theTiMo12Zr6Fe2 alloy for orthopaedic implants: in vitro biocompatibility studyby using primary human fibroblasts and osteoblasts, Biomaterials 23 (2002)2863–2869.

29] D.M. Gordin, T. Gloriant, G. Texier, I. Thibon, D. Ansel, J.L. Duval, M.D. Nagel,Development of a �-type Ti-12Mo-5Ta alloy for biomedical applications: cyto-compatibility and metallurgical aspects, Journal of Materials Science: Materialsin Medicine 15 (2004) 885–891.

30] S. Nag, R. Banerjee, H.L. Fraser, Microstructural evolution and strengtheningmechanisms in Ti-Nb-Zr-Ta, Ti-Mo-Zr-Fe and Ti-15Mo biocompatible alloys,Materials Science and Engineering C 25 (2005) 357–362.

31] I.J. Polmear, Light Alloys From Traditional Alloys to Nanocrystals, 4th ed., Else-vier, 2006, pp. 301.

32] J.W. Lu, Y.Q. Zhao, P. Ge, H.Z. Niu, Microstructure and beta grain growth behaviorof Ti–Mo alloys solution treated, Materials Characterization 84 (2013) 105–111.

33] Y.L. Zhou, D.M. Luo, Microstructures and mechanical properties of Ti–Mo alloyscold-rolled and heat treated, Materials Characterization 62 (2011) 931–937.

34] N.T.C. Oliveira, A.C. Guastaldi, Electrochemical behavior of Ti–Mo alloys appliedas biomaterial, Corrosion Science 50 (2008) 938–945.

35] N.T.C. Oliveira, G. Aleixo, R. Caram, A.C. Guastaldi, Development of Ti–Mo alloysfor biomedical applications: Microstructure and electrochemical characteriza-tion, Materials Science and Engineering A 452 (453) (2007) 727–731.

36] H.Y. Kim, Y. Ikehara, J.I. Kim, H. Hosoda, S. Miyazaki, Martensitic transformation,shape memory effect and superelasticity of Ti–Nb binary alloys, Acta Materialia54 (2006) 2419–2429.

37] H. Tobe, H.Y. Kim, T. Inamura, H. Hosoda, T.H. Nam, S. Miyazaki, Effect of Nbcontent on deformation behavior and shape memory properties of Ti–Nb alloys,Journal of Alloys and Compounds 577S (2013) S435–S438.

38] Y.B. Wang, Y.F. Zheng, Corrosion behaviour and biocompatibility evaluation oflow modulus Ti–16Nb shape memory alloy as potential biomaterial, MaterialsLetters 63 (2009) 1293–1295.

39] Y.B. Wang, Y.F. Zheng, The microstructure and shape memory effect of Ti–16at.% Nb alloy, Materials Letters 62 (2008) 269–272.

imica

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

R. Chelariu et al. / Electroch

40] L.H. de Almeida, I.N. Bastos, I.D. Santos, A.J.B. Dutra, C.A. Nunes, S.B.Gabriel, Corrosion resistance of aged Ti–Mo–Nb alloys for biomedical appli-cations, Journal of Alloys and Compounds (2014), http://dx.doi.org/10.1016/j.jallcom.2014.01.173.

41] L. Xu, S. Xiao, J. Tian, Y. Chen, Microstructure, mechanical properties anddry wear resistance of ˇ-type Ti - 15Mo - xNb alloys for biomedical appli-cations, Transactions of the Nonferrous Metals Society of China 23 (2013)692–698.

42] Y. Al-Zain, H.Y. Kim, H. Hosoda, T.H. Nam, S. Miyazaki, Shape memory propertiesof Ti–Nb–Mo biomedical alloys, Acta Materialia 58 (2010) 4212–4223.

43] C.M. Lee, C.P. Ju, J.H. Chern Lin, Structure-property relationship of cast Ti-Nballoys, Journal of Oral Rehabilitation 29 (2002) 314–322.

44] D. Mantovani, Shape memory alloys: Properties and biomedical applications,Journal of Minerals, Metals and Materials Society 52–10 (2000) 36–44.

45] S. Shabalovskaya, On the nature of biocompatibility and medical applications ofNiTi shape memory alloys, Biomedical Materials Engineering 6 (1996) 267–290.

46] M. Balazic, J. Kopac, M.J. Jackson, W. Ahmed, Review: titanium and titaniumalloy applications in medicine, International Journal of Nano and Biomaterials1 (2007) 3–34.

47] M. Geetha, A.K. Singh, R. Asokamani, A.K. Gogia, Ti based biomaterials, the ulti-mate choice for orthopaedic implants–a review, Progress in Material Science54 (2009) 397–425.

48] H.Y. Kim, H. Satoru, J.I. Kim, H. Hosoda, S. Miyazaki, Mechanical propertiesand shape memory behavior of Ti-Nb alloys, Materials Transactions 45 (2004)2443–2448.

49] S.B. Gabriel, L.H. de Almeida, C.A. Nunes, J. Dille, G.A. Soares, Maximisationof the ration of microhardness to the young’s modulus of Ti-12Mo-13Nballoy through microstructure changes, Materials Science and Engineering C 33(2013) 3319–3324.

50] E.W. Collings, Physical Metallurgy of Titanium Alloys, in: R. Boyer, G. Welsch,E.W. Collings (Eds.), Materials Properties Handbook: Titanium Alloys, ASM,International, Materials Park, OH, 1994, p. 10.

51] K.J. Bundy, Corrosion Testing in In Vivo Environments, in: R. Baboian (Ed.),Corrosion Tests and Standards: Application and Interpretation, ASTM, Philadel-phia, PA, 2005, p. 500.

52] D. Mareci, I. Rusu, R. Chelariu, G. Bolat, C. Munteanu, D. Sutiman, R.M. Souto,Application of dynamic electrochemical impedance spectroscopy to the eval-uation of the corrosion resistance of a historic bronze object in artificial acidrainwater, European Journal of Science and Theology 9 (2013) 189–199.

53] G. Airoldi, G. Riva, M. Vanelli, V. Filippi, G. Garanttini, Oral environmenttemperature changes induced by cold/hot liquid intake, American Journal ofOrthodontal and Dentofacial Orthopaedics 112 (1997) 58–63.

54] G. Ciurescu, J. Izquierdo, J.J. Santana, D. Mareci, D. Sutiman, S. González, R.M.Souto, Characterization of the localized surface chemical activity of Ti-Moand Ti-Ta alloys for biomedical applications using scanning electrochemi-cal microscopy, International Journal of Electrochemical Science 7 (2012)7404–7424.

55] M.J. Blackburn, J.C. Williams, Phase Transformations in Ti-Mo and Ti-V alloys,Transactions of The Metallurgical Society of the American Institute of Mining,Metallurgical and Petroleum Engineers 242 (1968) 2461–2469.

56] E. Sukedai, H. Hashimoto, M. Tomita, Investigation of omega-phase in Ti–Moalloys by high resolution electron microscopy, image processing and dark-fieldmethods, Philosophical Magazine A 64 (1991) 1201–1208.

57] F. Prima, P. Vermaut, G. Texier, D. Ansel, T. Gloriant, Evidence ofalpha-nanophase heterogeneous nucleation from omega particles in a beta-metastable Ti-based alloy by high-resolution electron microscopy, ScriptaMaterialia 54 (2006) 645–648.

58] W.S. Tait, An Introduction to Electrochemical Corrosion Testing for PracticingEngineers and Scientists, PairODocs Publications, Racine, WI, 1994, pp. 37.

59] E. Alkhateeb, S. Virtanen, Influence of surface self-modification in Ringer’s

solution on the passive behavior of titanium, Journal of Biomedical MaterialsResearch A 75 (2005) 934–940.

60] I. Milosev, B. Kapun, The corrosion resistance of Nitinol alloy in simulated phys-iological solutions. Part 1: The effect of surface preparation, Materials Scienceand Engineering C 32 (2012) 1087–1096.

[

Acta 137 (2014) 280–289 289

61] B.L. Wang, Y.F. Zheng, L.C. Zhao, Effects of Hf content and immersion timeon electrochemical behavior of biomedical Ti-22Nb-xHf alloys in 0.9% NaClsolution, Materials and Corrosion 60 (2009) 330–335.

62] G. Bolat, J. Izquierdo, J.J. Santana, D. Mareci, R.M. Souto, Electrochemical char-acterization of ZrTi alloys for biomedical applications, Electrochimica Acta 88(2013) 447–456.

63] T.P. Hoar, D.C. Mears, Corrosion-resistant alloys in chloride solutions: materialsfor surgical implants, Proceedings of the Royal Society London A 294 (1966)486–510.

64] Y.J. Bai, Y.B. Wang, Y. Cheng, F. Deng, Y.F. Zheng, S.C. Wei, Comparative study onthe corrosion behavior of Ti–Nb and TMA alloys for dental application in variousartificial solutions, Materials Science and Engineering C 31 (2011) 702–711.

65] M. Atapour, A.L. Pilchak, G.S. Frankel, J.C. Williams, Corrosion behavior of �titanium alloys for biomedical applications, Materials Science and EngineeringC 31 (2011) 885–891.

66] Y.F. Zheng, B.L. Wang, J.G. Wang, C. Li, L.C. Zhao, Corrosion behaviour ofTi–Nb–Sn shape memory alloys in different simulated body solutions, MaterialsScience and Engineering A 438–440 (2006) 891–895.

67] B.L. Wang, Y.F. Zheng, L.C. Zhao, Effects of Hf content and immersion timeon electrochemical behavior of biomedical Ti-22Nb-xHf alloys in 0.9% NaClsolution, Materials and Corrosion 60 (2009) 330–335.

68] X. Cheng, S.G. Roscoe, Corrosion behavior of titanium in the presence of calciumphosphate and serum proteins, Biomaterials 26 (2005) 7350–7356.

69] S. Tamilselvi, N. Rajendran, Electrochemical studies on the stability and corro-sion resistance of Ti-5Al-2Nb-1Ta alloy for biomedical applications, Trends inBiomaterials & Artificial Organs 20 (2006) 49–52.

70] S. Tamilselvi, R. Murugaraj, N. Rajendran, Electrochemical impedance spec-troscopic studies of titanium and its alloys in saline medium, Materials andCorrosion 58 (2007) 113–120.

71] S. Tamilselvi, N. Rajendran, In vitro corrosion behaviour of Ti-5Al-2Nb-1Ta alloyin Hanks solution, Materials and Corrosion 58 (2007) 285–289.

72] S.E. Kim, J.H. Son, Y.T. Hyun, H.W. Jeong, Y.T. Lee, J.S. Song, J.H. Lee, Electrochem-ical corrosion of novel beta titanium alloys, Metals and Materials International13 (2007) 151–154.

73] M.E.P. Souza, L. Lima, C.R.P. Lima, C.A.C. Zavaglia, C.M.A. Freire, Effects of pHon the electrochemical behavior of titanium alloys for implant applications,Journal of Materials Science: Materials in Medicine 20 (2009) 549–552.

74] C.W. Chan, H,C, Man, T.M. Yue, Effect of post-weld heat-treatment on the oxidefilm and corrosion behaviour of laser-welded shape memory NiTi wires, Cor-rosion Science 56 (2012) 158-167.

75] J. Pan, D. Thierry, C. Leygraf, Electrochemical impedance spectroscopy study ofthe passive oxide film on titanium for implant application, Electrochimica Acta41 (1996) 1143–1153.

76] V.A. Alves, R.Q. Reis, I.C.B. Santos, D.G. Souza, T. de, F. Gonc alves, M.A. Pereira-da-Silva, A. Rossi, L.A. da Silva, In situ impedance spectroscopy study of theelectrochemical corrosion of Ti and Ti-6Al-4 V in simulated body fluid at 25 ◦Cand 37 ◦C, Corrosion Science 51 (2009) 2473–2482.

77] D. Wallinder, J. Pan, C. Leygraf, A. Delblanc-Bauer, EIS and XPS study of surfacemodification of 316LVM stainless steel after passivation, Corrosion Science 41(1999) 275–289.

78] F. Mansfeld, Analysis and interpretation of EIS data for metals and alloys, Tech-nical Report 26, Solartron-Schlumberger, Farnborough, 1993, p. 4. 1-4.24.

79] G. Bolat, D. Mareci, R. Chelariu, J. Izquierdo, S. González, R.M. Souto, Investiga-tion of the electrochemical behaviour of TiMo alloys in simulated physiologicalsolutions, Electrochimica Acta 113 (2013) 470–480.

80] I.C. Lavos-Valereto, S. Wolynec, I. Ramires, A.C. Guastaldi, I. Costa, Electrochemi-cal impedance spectroscopy characterization of passive film formed on implantTi-6Al-7Nb alloy in Hank’s solution, Journal of Materials Science: Materials inMedicine 15 (2004) 55–59.

81] A.K. Shukla, R. Balasubramaniam, S. Bhargava, Effect of replacement of V by Fe

and Nb on passive film behavior of Ti–6Al–4V in simulated body fluid condi-tions, Journal of Alloys and Compounds 389 (2005) 144–152.

82] S. Tamilselvi, V. Raman, N. Rajendran, Corrosion behaviour of Ti–6Al–7Nb andTi–6Al–4V ELI alloys in the simulated body fluid solution by electrochemicalimpedance spectroscopy, Electrochimica Acta 52 (2006) 839–846.