You are on page 1of 7

Available online at www.sciencedirect.

com

Acta Biomaterialia 5 (2009) 399–405


www.elsevier.com/locate/actabiomat

Electrochemical stability and corrosion resistance of Ti–Mo alloys


for biomedical applications
N.T.C. Oliveira *, A.C. Guastaldi
Grupo de Biomateriais, Departamento de Fı́sico-Quı́mica, Instituto de Quı́mica, Universidade Estadual Paulista-UNESP, Rua Francisco Degni, s/n,
P.O. Box 355, CEP 14801-970, Araraquara – SP, Brazil

Received 13 February 2008; received in revised form 11 June 2008; accepted 15 July 2008
Available online 25 July 2008

Abstract

Electrochemical behavior of pure Ti and Ti–Mo alloys (6–20 wt.% Mo) was investigated as a function of immersion time in electrolyte
simulating physiological media. Open-circuit potential values indicated that all Ti–Mo alloys studied and pure Ti undergo spontaneous
passivation due to spontaneously formed oxide film passivating the metallic surface, in the chloride-containing solution. It also indicated
that the addition of Mo to pure Ti up to 15 wt.% seems to improve the protection characteristics of its spontaneous oxides. Electrochem-
ical impedance spectroscopy (EIS) studies showed high impedance values for all samples, increasing with immersion time, indicating an
improvement in corrosion resistance of the spontaneous oxide film. The fit obtained suggests a single passive film present on the metals’
surface, improving their resistance with immersion time, presenting the highest values to Ti–15Mo alloy. Potentiodynamic polarization
showed a typical valve-metal behavior, with anodic formation of barrier-type oxide films, without pitting corrosion, even in chloride-
containing solution. In all cases, the passive current values were quite small, and decrease after 360 h of immersion. All these electro-
chemical results suggest that the Ti–15Mo alloy is a promising material for orthopedic devices, since electrochemical stability is directly
associated with biocompatibility and is a necessary condition for applying a material as biomaterial.
Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Ti–Mo alloys; Corrosion resistance; Biocompatible alloys; Titanium alloys

1. Introduction sion and wear products in the tissues surrounding the


implant may result in a cascade of events leading to peri-
Titanium and its alloys have been used extensively in the prosthetic bone loss [3]. Hence, protection of an implant
last several decades as materials for orthopedic implants, material from corrosion is indispensable [3].
dental implants, and medical devices [1]. This is due to Growing interest has been observed recently in the
the formation of a passive film on their surface, consisting development of a new generation of biocompatible, corro-
mainly of amorphous titanium dioxide (TiO2), which is sion and wear-resistant materials [4]. Attempts were made
responsible for both their corrosion resistance in several to develop titanium alloys of different compositions to
media including the human body environment, and their achieve better performance in terms of biomechanical com-
biocompatibility [2,3]. patibility (by reducing the Young’s modulus) and biochem-
Implant corrosion caused by the reaction with body flu- ical compatibility (by excluding non-toxic elements) [3].
ids and tissues seems to affect the fatigue life and ultimate Research on many titanium biomaterials is done by
strength of the material, leading to mechanical failure of focusing on b type titanium alloys, because processing vari-
the implants [3]. Further, the presence of particulate corro- ables can be controlled to produce selected results.
Enhanced properties such as lower modulus of elasticity,
*
Corresponding author. Tel.: +55 16 33016655; fax: +55 16 33227932. increased corrosion resistance, and improved tissue
E-mail address: ntco@yahoo.com (N.T.C. Oliveira). response are possible when compared with (a + b) type

1742-7061/$ - see front matter Ó 2008 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actbio.2008.07.010
400 N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405

alloys [5]. Therefore, b type titanium alloys composed of ducibility of Eoc results, this analysis was repeated four
non-toxic elements such as Nb, Ta, Zr, Mo, and Sn show- times measured at the same sample, after new polish of
ing lower modulus of elasticity and greater strength should the surface’s alloys, for all samples.
be developed [6].
In recent years, Ti–Mo alloys employed as biomaterials 2.3. Electrochemical impedance spectroscopic studies
have been studied with emphasis on their microstructure
and mechanical properties [7]. Ho et al. [8], Sugano et al. Electrochemical impedance spectroscopy (EIS) measure-
[9], Guo and Enomoto [10], Sukedai et al. [11], Liu et al. ments were carried out at open-circuit potential using a
[12], and Oliveira et al. [7], for example, have conducted Solartron 1260 electrochemical frequency response ana-
different studies on phase transformations, stress release, lyzer (FRA) system. The impedance spectra were acquired
and mechanical properties of different Ti–Mo alloys, but in the frequency range from 100 kHz to 10 mHz with a per-
there are only a few studies about their electrochemical turbation signal of 10 mV. EIS plots were obtained after
behavior [13–16]. the specimens were immersed in the test solution for differ-
In the present work, the electrochemical behavior of ent numbers of hours (1, 24, 48, 72, 168, 240, and 360 h).
Ti–Mo alloys applied as biomaterials was investigated as Sample surface morphology was studied by SEM, after
a function of immersion time (up to 360 h) in simulated 0, 1 and 360 h of immersion.
physiological media. Previously, Oliveira et al. [7,16] inves-
tigated Ti–Mo alloys just with 1 h of immersion in the same 2.4. Potentiodynamic polarization studies
solution.
Potentiodynamic polarization studies were carried out
2. Experimental after 1 and 360 h of immersion in Ringer’s solution for
pure titanium and Ti–Mo alloys. Potentiodynamic polari-
2.1. Materials and solutions zation scans were carried out at a scan rate of 1 mV s 1
in the range from –0.8 to 5.0 V vs. SCE using a Solartron
Ti–Mo alloys with different compositions (6, 10, 15, and 1287 potentiostat/galvanostat.
20 wt.% of Mo), were prepared by the authors using an arc-
melting furnace under an ultra-pure argon atmosphere, fol- 3. Results and discussion
lowing a procedure described in the literature [7,17]. These
alloys were characterized using energy-dispersive X-ray To characterize the electrochemical behavior of the Ti–
analysis (EDX), X-ray fluorescence (XRF), scanning elec- Mo alloys in a solution that simulated the physiological
tron microscopy (SEM), chemical mapping, and X-ray dif- media, electrochemical studies were conducted on the
fraction (XRD); these data were presented in previous alloys as a function of immersion time. For the sake of
papers [7,16]. The XRD analysis [7] showed that the crystal comparison, experiments were also obtained for pure Ti.
structure of the Ti–Mo alloys is sensitive to the Mo concen- The chloride-containing solution used was Ringer’s saline,
tration: Ti–6Mo (a type), Ti–10Mo (a + b type), Ti–15Mo, which is the usual electrolyte employed to simulate the
and Ti–20Mo (b type) alloys. physiological media [7,16–21].
Electrochemical experiments were conducted in a stan-
dard three-electrode cell with 0.44 cm2 of exposed area in 3.1. Open-circuit potential analyses
the working electrode, having a platinum mesh as a counter
electrode and a saturated calomel electrode (SCE) as refer- The time profiles of the open-circuit potential obtained
ence. Working electrolyte was a naturally aerated aqueous for the samples were quite similar (Fig. 1). As can be seen,
Ringer’s physiological solution (NaCl 8.61 g l 1, CaCl2 Eoc changes quickly towards more positive potentials dur-
0.49 g l 1, KCl 0.30 g l 1). Prior to any immersion, the ing the first hour. After that, Eoc changes more slowly until
working electrodes were polished with 1500 grade silicon it reaches a quasi-stationary value at 72 h, not changing
carbide paper and rinsed with distilled and deionized significantly after that. This fact indicates that all Ti–Mo
(Milli-QÒ) water. All experiments were carried out at alloys studied and pure Ti undergo spontaneous passiv-
37 °C, and the Ti–Mo alloys were studied on as-cast condi- ation due to spontaneously formed oxide film passivating
tions, at different immersion times. the metallic surface, in the chloride-containing solution.
Similar behavior was found by Cai et al. [1], where an ini-
2.2. Open-circuit potential analyses tial increase in the Eocs during the early hours followed by
stabilization (–205 to –85 mV vs. SCE) observed on all the
Open-circuit potential measurements, Eoc, for each alloy specimens (CP TI, Ti–6Al4 V, Ti–6Al–7Nb and Ti–13Nb–
were carried out on freshly polished samples, in naturally 13Zr) suggests that a protective passive film formed rapidly
aerated aqueous electrolyte without stirring, immediately on the metal surfaces in the artificial saliva and remained
after polishing. The Eoc was continuously monitored dur- stable during the entire immersion period (6  104 s) [1].
ing the first hour, and then its values were recorded after By comparing the Eoc values (Fig. 1), it can be observed
24, 48, 72, 168, 240, and 360 h. In order to verify the repro- that the Ti–15Mo alloy had the most positive values, while
N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405 401

3.2. Electrochemical impedance spectroscopic studies

Impedance spectra for pure Ti and Ti–Mo alloys at dif-


ferent immersion hours (1, 24, 48, 72, 168, 240, and 360 h)
in Ringer’s solution are represented as Bode plots, exempli-
fied in Fig. 2 for pure Ti, Ti–6Mo, and Ti–15Mo.
These plots exhibit low impedance values at 1 h of
immersion, whereas impedance values increased with
immersion time. However, no significant change in their
behavior was observed after 24 h. This fact could be due
to a complete coverage of the surface of newly formed
spontaneous oxide layers [2].

Fig. 1. Open-circuit potential vs. time profile for pure Ti and Ti–Mo
alloys with 6, 10, 15, and 20Mo wt.% after different immersion times in
Ringer physiological solution.

the Ti–6Mo and Ti–10Mo alloys, as well as pure Ti, had


similar values, and finally, the Ti–20Mo alloy showed the
most negative values. These results indicate that the addi-
tion of Mo to pure Ti up to 15 wt.% seems to improve
the protection characteristics of its spontaneous oxides.
However, at higher Mo concentration values, the effect is
inverted, making the Eoc values for the Ti–20Mo alloy
more negative, i.e., less noble, than for the other materials
studied.
There are only a few studies about electrochemical
behavior of Ti–Mo alloys [13–16], but it is well know that
b titanium alloys generally exhibit superior corrosion resis-
tance in comparison to a + b alloys [3]. However, their cor-
rosion resistance is also found to depend on several factors
such as composition, environment and microstructure [3].
As pointed out by Geetha et al. [3], the effect of the compo-
sition can be understood from the fact that the corrosion
resistance of a b alloy such as Ti–5Mo–5Zr–3Al is higher
when compared to pure Ti, and the influence of microstruc-
ture on corrosion is understood from the studies carried
out by Yu and Scully [26] on b ST Ti–15Mo–3Nb–3Al.
From their studies, it is observed that the b solution-treated
microstructure exhibits superior corrosion performance
when compared to the aged one. The decrease in the resis-
tance of the aged sample is due to the partitioning of the
alloying elements occurring during the aging process [3].
Thus, the corrosion studies carried out on various alloys
clearly indicate that the composition and microstructural
features decide the corrosion characteristics of the alloy
in addition to the texture and crystal structure. As men-
tioned before in the present paper, Oliveira et al. [7] found
that the crystal structure of the Ti–Mo alloys is sensitive to
the Mo concentration: Ti–6Mo (a type), Ti–10Mo (a + b
type), Ti–15Mo, and Ti–20Mo (b type) alloys; so this can
be a possible reason of the improvement of corrosion resis-
tance for Ti–15Mo alloy in comparison to other Fig. 2. Bode diagrams for (a) pure Ti, (b) Ti–6Mo, and (c) Ti–15Mo
compositions. alloys, after different immersion times in Ringer physiological solution.
402 N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405

The phase angle, h, is a sensitive parameter used to indi-


cate the presence of additional time constants in the imped-
ance spectra at the highest and lowest frequencies. The use
of the log |Z| vs. log f format enables an equal representa-
tion of all experimental data over the entire frequency
domain [22].
The Bode plots in Fig. 2 in the frequency range between
100 kHz and 10 mHz present only one time constant and a
near-capacitive response with a phase angle close to 80°
and linear variation between log |Z| and log f, at a wide
range of frequencies, with a slope close to 1. The phase
angle observed at low frequency for immediate immersion
was near 40. However, at intermediate frequencies, the
phase angle shifted to 80° and remained constant over a
wide range of frequencies. After 24 h of immersion, the
phase angle remained near 80° even at low frequencies,
indicating a near-capacitive response for all metals studied.
High impedance values (in the order of 106 X cm2) were
obtained for medium to low frequencies in all samples, sug-
gesting high corrosion resistance in the electrolytes used,
and are indicative of a single, thin passive oxide film pres-
ent on the surface of the samples [2] since the beginning of
immersion. The increase in impedance values with immer-
sion time indicated an improvement in corrosion resistance
of the spontaneous oxide film in the chloride-containing
solution, for pure Ti and all Ti–Mo alloys studied.
SEM analyses were carried out on the samples’ surface
to characterize possible changes, in addition to forms and
extent of corrosion after 0, 1, and 360 h. Exemplifying Fig. 3. SEM micrographs for pure Ti, Ti–6Mo, and Ti–15Mo alloys after
0, 1, and 360 h of immersion in Ringer physiological solution.
the results obtained for all alloys, in Fig. 3 are shown the
pure Ti, Ti–6Mo, and Ti–15Mo micrographs; as can be
seen, there were no significant morphological changes constant phase element (CPE), which takes into account
between the various immersion times, i.e., no pitting, the capacitive behavior of the film. Gonzalez and Mirza-
cracks, or other defects appeared on the metal surfaces Rosca [20] proposed Rs( QRp) as the equivalent circuit
until 360 h of immersion, even in presence of Cl ions, con- model to fit the EIS data in the case of a single passive film
firming the electrochemical results obtained. present on the metal surface. This equivalent circuit has
been generally used to fit oxides grown on Ti alloys under
3.2.1. Selection of the equivalent circuit different situations [2,17,22,23].
Shukla and Balasubramaniam [23], using potentiody- It can be seen in Table 1 and Fig. 6a that the capacitance
namic polarization and EIS techniques, studied the effect decrease, in almost all cases, with the immersion time and n
of immersion time in a solution simulating the physiological values are near 1, so Q presents an almost pure capacitive
media, Hank’s solution, for Ti–6Al–4V and Ti–13Nb–13Zr behavior. The film resistance, RP (Table 1 and Fig. 6b) has
on their electrochemical behavior. The potentiodynamic the same order of magnitude, 106 X cm 2, obtained by
polarization and EIS studies revealed that the passive film other authors studding different Ti alloys [17,22,23], and
behavior of untreated Ti-alloys did not change significantly it increases with immersion time, demonstrating that the
with immersion time in Hank’s solution, and data could be spontaneously formed oxide film on the alloys and on pure
modeled assuming a single passive film to be present. On the Ti becomes more resistive with immersion time. Finally,
other hand, the EIS results obtained by Tamilselvi et al. [2] the RP values for Ti–15Mo showed that this alloy has bet-
for Ti–6Al–7Nb and Ti–6Al–4V ELI alloys indicated the ter corrosion resistance and higher stability than the other
presence of a single passive layer immediately after immer- metals studied at 360 h of immersion in Ringer’s solution,
sion for both alloys. However, the alloys immersed for which agrees with and confirms the results obtained in
360 h in SBF solution showed the presence of a bi-layered the open circuit potential studies (Fig. 1).
surface corresponding to an inner layer and an outer layer.
At the present paper, a satisfactory fit (see Fig. 4, for Ti– 3.3. Potentiodynamic polarization studies
15Mo) of all data could be obtained using a simple
Rs(QRp) circuit (Fig. 5), where Rs and Rp are the solution To find more appropriated metals or alloys to be applied
and the parallel (film) resistances, respectively, and Q is a as biomaterials is usual to analyze their in vitro corrosion
N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405 403

Table 1
Values of the circuit parameters derived from the fitting of the impedance
data measured at open circuit potential for passive films grown sponta-
neously on pure Ti and Ti–Mo alloys with 6, 10, 15, and 20Mo wt.%, using
the equivalent electrical circuit RS(QRP), from Fig. 5
t (h) Rs (X cm 2) Rp (X cm 2) n Q (F cm 2
sn 1)
Ti–6Mo
1 76.42 2.27  105 0.933 4.84  10 5

24 74.21 1.40  106 0.939 3.35  10 5

48 74.6 1.61  106 0.948 2.97  10 5

72 78.7 1.66  106 0.946 2.93  10 5

168 71.1 1.78  106 0.945 2.91  10 5

240 74.1 1.98  106 0.949 2.90  10 5

360 72.9 2.26  106 0.939 2.89  10 5

Ti–10Mo
1 64.7 2.78  105 0.906 6.45  10 5

24 57.0 6.96  105 0.932 3.52  10 5

48 63.4 1.00  106 0.931 3.32  10 5

72 64.0 1.47  106 0.939 3.22  10 5

168 59.9 1.66  106 0.935 3.15  10 5

240 61.4 2.12  106 0.932 3.11  10 5

360 61.0 2.19  106 0.932 3.09  10 5

Ti–15Mo
1 68.5 1.70  105 0.904 7.98  10 5

24 57.1 9.29  105 0.943 2.82  10 5

48 58.4 1.80  106 0.947 2.64  10 5

72 57.8 1.91  106 0.957 2.44  10 5

168 56.4 2.24  106 0.953 2.54  10 5

240 56.1 2.47  106 0.959 2.49  10 5

360 56.8 2.80  106 0.951 2.54  10 5

Fig. 4. (a) Nyquist and (b) Bode diagrams, for Ti–15Mo alloy after 360 h Ti–20Mo
of immersion in Ringer physiological solution and its fitting with the 1 65.6 8.70  104 0.882 1.33  10 4

equivalent circuit, Rs(RpQ), from Fig. 5. 24 71.2 4.87  105 0.921 6.54  10 5

48 72.2 1.12  106 0.923 6.67  10 5

72 73.6 1.79  106 0.94 6.21  10 5

168 70.2 1.89  106 0.932 6.72  10 5

240 66.3 1.90  106 0.931 6.73  10 5

360 70.3 1.89  106 0.927 6.72  10 5

Ti
1 57.4 1.66  105 0.899 6.96  10 5

24 64.1 1.23  106 0.896 5.73  10 5

48 64.9 1.48  106 0.908 5.63  10 5

72 54.4 1.59  106 0.926 5.41  10 5

168 64.1 1.87  106 0.923 5.39  10 5

240 65.4 1.90  106 0.922 5.36  10 5


5
360 60.0 2.30  106 0.945 5.32  10

resistance in harder and more general conditions, usually


up to very high potentials [7,16–19,24–30], since electro-
chemical stability is directly associated with biocompatibil-
ity and is a necessary condition for applying a material as
biomaterial.
The corrosion resistance of the Ti–Mo alloys has
already been studied [7,16] after 1 h of immersion in neutral
(Na2SO4) solution or simulating the physiological media
(Ringer solution). But with the purpose of investigating
the influence of higher immersion times on corrosion resis-
tance of these alloys in Ringer solution, potentiodynamic
polarization scanning at 1 mV s 1 was carried out after 1
and 360 h of immersion in this solution, in the potential
Fig. 5. Equivalent circuit, Rs(RpQ), used for fitting the experimental data. range from –0.8 to 5.0 V (vs. SCE). Exemplifying all the
404 N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405

Fig. 7. Potentiodynamic polarization for Ti–15Mo alloy running after 1


and 360 h of immersion in Ringer physiological solution. v = 1 mV s 1.

Table 2
Passive current, ipas, for pure Ti and Ti–Mo alloys with 6, 10, 15, and 20
Mo wt.% in Ringer solution (v = 1 mV s 1)
Time (h) ipas (lA cm 2)
Sample
Ti Ti–6Mo Ti–10Mo Ti–15Mo Ti–20Mo
01 8.5 ± 0.3 7.0 ± 0.6 9.1 ± 0.4 10.3 ± 0.5 13.2 ± 0.8
360 6.1 ± 0.5 4.5 ± 0.4 3.8 ± 0.6 4.1 ± 0.7 7.8 ± 0.6

NaOH-treated Ti–6Al–4V and Ti–13Nb–13Zr after 168 h


of immersion in Hank’s solution revealed a potential range
in which the passive current density was lower than the
Fig. 6. (a) Capacitance, Q, and (b) polarization resistance, RP, for pure Ti passive current density after 1 h immersion. This feature
and Ti–Mo alloys with 6, 10, 15, and 20Mo wt.% after different immersion was not observed in the case of CP Ti. The passive current
times in Ringer physiological solution. densities for all the three NaOH-treated alloys (CP Ti, Ti–
6Al–4V, and Ti–13Nb–13Zr) showed a slight decrease with
results, Fig. 7 presents the voltammograms obtained for immersion time in Hank’s solution, which can be attributed
Ti–15Mo alloy after 1 and 360 h of immersion. to changes in surface nature with time [23].
In the potentiodynamic polarizations, Fig. 7, it can be These results confirm the evidence presented above that
noted that even in the solution containing Cl ions, they the addition of Mo to pure Ti increases its corrosion resis-
showed typical valve-metal behavior [17], i.e., a region of tance, in accordance with studies by Sakaguchi et al. [13],
formation and growth of the anodic oxide in the passive in which corrosion resistance of Ti–30Mo in 35% HCl stud-
region, without transpassivation (pitting corrosion) until ied by electrochemical analysis showed that this alloy is
the final growth potential of 5.0 V (SCE), indicating high more resistant than pure Ti, due to lower current density
corrosion resistance for these metals in the solutions that of the alloy.
simulated the physiological media. The valve-metal profile
presented by these alloys was described in detail in previous 4. Conclusions
papers [7,16].
In all cases, the passive current values, ipas, correspond- The electrochemical behavior of pure Ti and Ti–Mo
ing to the passive region, i.e., the second plateau, were alloys (6 to 20Mo wt.%) applied as biomaterials was inves-
quite small (Table 2), having the same order of magnitude, tigated by EIS, open-circuit potential measurements, and
near 10 lA cm 2, indicating that the oxide films formed potentiodynamic polarization, as a function of immersion
under the studied conditions gave good protection to the time in a electrolyte simulating the physiological media.
metal substrate. It can be seen that the ipas values decrease The results can be summarized as follows:
after 360 h of immersion (Table 2 and Fig. 7), suggesting
that the spontaneous oxide film formed on these samples 1. Open-circuit potential values indicated that all Ti–Mo
improves its corrosion resistance with the immersion, even alloys studied and pure Ti undergo spontaneous passiv-
in chloride-containing solution. The polarization curves ation due to spontaneously formed oxide film passivat-
obtained by Shukla and Balasubramaniam [23] of ing the metallic surface, in the chloride-containing
N.T.C. Oliveira, A.C. Guastaldi / Acta Biomaterialia 5 (2009) 399–405 405

solution. Ti–15Mo had the highest Eoc values, while Ti– [9] Sugano M, Tsuchida Y, Satake T, Ikeda M. A microstructural study
20Mo had the lowest values, indicating that the addition of fatigue fracture in titanium–molybdenum alloys. Mater Sci Eng A
1998;1–2:163–8.
of Mo to pure Ti up to 15 wt.% seems to improve the [10] Guo H, Enomoto M. Surface reconstruction associated with a
protection characteristics of its spontaneous oxides. precipitation in a Ti–Mo alloy. Scripta Mater 2006;54:1409–13.
2. High impedance values were obtained for all samples, [11] Sukedai E, Yoshimitsu D, Matsumoto H, Hashimoto H, Kiritani. b
and its increase with immersion time indicated an to x transformation due to aging in a Ti–Mo alloy deformed in
improvement in corrosion resistance of the spontaneous impact compression. Met Mater Sci Eng A 2003;1–3:133–8.
[12] Liu Y, Wei WF, Zhou KC, Chenl F, Tang HP. Microstructures and
oxide film. The fit obtained suggests a single passive film mechanical behavior of PM Ti–Mo alloy. J Central South Univ
present on the metals’ surface, with resistance improving Technol 2003;2(10):81–6.
with immersion time. The Ti–15Mo alloy had the highest [13] Sakaguchi S, Nakahara K, Hayashi Y. Development of sintered Ti–
RP values, showing that the alloy with 15 wt.% had better 30mass% Mo alloy and its corrosion properties. Met Mater Int
corrosion resistance than the other Ti–Mo alloys studied. 1999;2(5):193–5.
[14] Habazaki H, Uozumi M, Konno H, Nagata S, Shimizu K. Formation
3. Potentiodynamic polarization showed a typical valve- of barrier-type amorphous anodic films on Ti–Mo alloys. Surf Coat
metal behavior, with anodic formation of barrier-type Technol 2003;169–170:151–4.
oxide films, without pitting corrosion even in a chlo- [15] Popa MV, Vasilescu E, Drob P, Vasilescu C. The modelling of passive
ride-containing solution. In all cases, the passive current film formation on a new titanium industrial alloy in very aggressive
values were quite small, and decrease after 360 h of media. Rev Chim 2005;56(9):908–12.
[16] Oliveira NTC, Guastaldi AC. Electrochemical behavior of Ti–Mo
immersion. alloys applied as biomaterial. Corros Sci 2008;50(4):938–45.
[17] Oliveira NTC, Biaggio SR, Piazza S, Sunseri C, DiQuarto F. Photo-
All these electrochemical results are connected and sug- electrochemical investigation of passive layers grown anodically on
gest that Ti–Mo alloys are promising materials for ortho- titanium alloys. Electrochim Acta 2004;49(26):4563–76.
pedic devices, because their electrochemical stability is [18] Oliveira NTC, Ferreira EA, Duarte LT, Biaggio SR, Rocha-Filho
RC, Bocchi N. Studies on the corrosion resistance of anodic oxides on
directly associated with biocompatibility, particularly in biocompatible alloys, Ti–50Zr e Ti–13Nb–13Zr. Electrochim Acta
the case of Ti–15Mo. 2006;51(10):2068–75.
[19] Oliveira NTC, Biaggio SR, Rocha-Filho RC, Bocchi N. Electro-
chemical studies on zirconium and its biocompatible alloys Ti–
Acknowledgements
50Zr at.% and Zr–2.5Nb wt.% in simulated physiologic media. J
Biomed Mater Res A 2005;74A(3):397–407.
The authors are grateful to FAPESP for scholarships [20] González JEG, Mirza-Rosca JC. Study of the corrosion behaviour of
(Proc. 04/11751-8) and Grants (Proc. 2005/04050-6) that titanium and some of its alloys for biomedical and dental implant
made this work possible. applications. J Electroanal Chem 1999;471:109–15.
[21] Schmidt H, Konetschny C, Fink U. Electrochemical behavior of ion
implanted Ti–6Al–4V in Ringer’s solution. Mater Sci Technol
References 1998;14:592–8.
[22] Metikos-Hukovic M, Kwokal A, Piljac J. The influence of niobium
[1] Cai Z, Shafer T, Watanabe I, Nunn ME, Okabe T. Electro- and vanadium on passivity of titanium-based implants in physiolog-
chemical characterization of cast titanium alloys. Biomaterials ical solution. Biomaterials 2003;24:3765–75.
2003;24:213–8. [23] Shukla AK, Balasubramaniam R. Effect of surface treatment on
[2] Tamilselvi S, Raman V, Rajendran N. Corrosion behaviour of Ti– electrochemical behavior of CP Ti, Ti–6Al–4V and Ti–13Nb–13Zr
6Al–7Nb and Ti–6Al–4V ELI alloys in the simulated body fluid alloys in simulated human body fluid. Corros Sci 2006;48:1696–720.
solution by electrochemical impedance spectroscopy. Electrochim [24] López MF, Gutiérrez A, Jiménez JA. In vitro corrosion behavior of
Acta 2006;52:839–46. titanium alloys without vanadium. Electrochim Acta 2002;47:1359–64.
[3] Geetha M, Kamachi Mudali U, Gogia AK, Asokamani R, [25] Blackwood DJ, Peter LM, Williams DE. Stability and open circuit
Baldev Raj. Influence of microstructure and alloying elements on breakdown of the passive oxide film on titanium. Electrochim Acta
corrosion behavior of Ti–13Nb–13Zr alloy. Corros Sci 1988;33:1143–9.
2004;46:877–92. [26] Yu SY, Scully JR. Corrosion and passivity of Ti–13Nb–13Zr in
[4] Czarnowska E, Wierzchón T, Maranda-Niedbala A. Properties of the comparison to other biomedical implant alloys. Corrosion
surface layers on titanium alloy and their biocompatibility in vitro 1997;53:965–76.
tests. J Mater Process Technol 1999;92–93:190–4. [27] Khan MA, Williams RL, Williams DF. In vitro corrosion and wear
[5] Kuroda D, Niinomi M, Masahiko M, Kato Y, Yashiro T. Design and of titanium alloys in the biological environment. Biomaterials
mechanical properties of new type titanium alloys for implants 1996;17:2117–26.
materials. J Biomed Mater Res 1995;29:943–50. [28] Khan MA, Williams RL, Williams DF. The corrosion behaviour of
[6] Niinomi M, Kuroda D, Fukunaga K, Morinaga M, Kato Y, Yashiro Ti–6Al–4V, Ti–6Al–7Nb and Ti–13Nb–13Zr in protein solutions.
T, et al. Corrosion wear fracture of new type biomedical titanium Biomaterials 1999;20:631–7.
alloys. Mater Sci Eng A 1999;263:193–9. [29] Khan MA, Williams RL, Williams DF. Conjoint corrosion and wear
[7] Oliveira NTC, Aleixo G, Caram R, Guastaldi AC. Electrochemical in titanium alloys. Biomaterials 1999;20:765–72.
behavior of Ti–Mo alloys applied as biomaterials. J Mater Sci Eng A [30] Lavos-Valereto IC, Wolynec S, Ramires I, Guastaldi AC, Costa I.
2007;452–453:727–31. Electrochemical impedance spectroscopy characterization of passive
[8] Ho WF, Ju CP, Chern Lin JH. Structure and properties of cast binary film formed on implant Ti–6Al–7Nb alloy in Hank’s solution. J
Ti–Mo alloys. Biomaterials 1999;20:2115–22. Mater Sci Mater Med 2004;25:55–9.

You might also like