Sei sulla pagina 1di 8

Available online at www.sciencedirect.

com

Electrochimica Acta 53 (2008) 48674874

Electrochemical corrosion behavior of a Ti35Nb alloy


for medical prostheses
Alessandra Cremasco, Wislei R. Osorio , Celia M.A. Freire,
Amauri Garcia, Rubens Caram
Department of Materials Engineering, State University of Campinas, UNICAMP, P.O. Box 6122,
13083-970 Campinas, SP, Brazil
Received 27 November 2007; received in revised form 1 February 2008; accepted 2 February 2008
Available online 12 February 2008

Abstract
Since the 1980s, the titanium alloys show attractive properties for biomedical applications where the most important factors are, firstly, biocompatibility, corrosion and mechanical resistances, low modulus of elasticity, very good strength to weight ratio, reasonable formability and
osseointegration. The aim of this study was to investigate the effects of two different heat treatments; furnace cooling and water quenching, on
the general electrochemical corrosion resistance of Ti35 wt%Nb alloy samples immersed in a 0.9% NaCl (0.15 mol L1 ) solution at 25 C and
neutral pH range. The samples were obtained using a non-consumable tungsten electrode furnace with a water-cooled copper hearth under argon
atmosphere. The microstructural pattern was examined by scanning electron microscopy (SEM) and X-ray diffractometry (XRD). In order to
evaluate the electrochemical corrosion behavior of such TiNb alloy samples, corrosion tests were performed by using electrochemical impedance
spectroscopy (EIS) and potentiodynamic polarization curves. Analyses of an equivalent circuit have also been used to provide quantitative support
for the discussions and understanding of the corrosion behavior. It was found that water quenching provides a microstructural pattern consisting of
an alpha-martensite acicular phase which decreases the material electrochemical performance due to the stress-induced martensitic transformation.
2008 Elsevier Ltd. All rights reserved.
Keywords: TiNb alloys; Biomaterials; Corrosion resistance; Microstructure; EIS diagrams

1. Introduction
Since the commercial introduction of Ti alloys in the early
1950s, these materials have in a relatively short time become
one of the backbone materials for aerospace, energy, chemical
industries, nuclear power plants and medical prostheses [1,2].
Since the 1980s, the titanium alloys are intensively applied
in the manufacturing of biomedical devices where the most
important factors are, firstly, biocompatibility, corrosion resistance, mechanical behavior and osseointegration (facility to
bone ingrowths or fixes in the metal implant with the patients
bone structure) [17]. The Ti6Al4V alloy was the first Ti alloy
registered as implant material in ASTM standards [8] and has
dominated wrought industry production. It has become a benchmark alloy against which others are compared. Its studies in vitro

Corresponding author. Tel.: +55 19 3521 3320; fax: +55 19 3289 3722.
E-mail address: wislei@fem.unicamp.br (W.R. Osorio).

0013-4686/$ see front matter 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.electacta.2008.02.011

and in vivo under human body environment are well reported in


the literature [911].
Recently, some studies have focused on the development of
vanadium and aluminum free Ti alloys since they exhibit high
cytotoxicity and negative tissue response in vivo and may induce
to senile dementia, neurological disorders and allergic reactions
[12,13]. In the last decade, Nb, Ta, Zr, Mo and Sn have been
selected as safe alloying elements to Ti alloys which are judged
to be non-toxic and non-allergic, and are considered as excellent
phase stabilizers elements [1316]. In this particular, Niobium,
a 4d transition metallic element, is one of the most effective titanium -stabilizer. In addition, atoms of Nb occupy exclusively
the Ti sites resulting in solidsolution strengthening and precipitation hardening [16]. The phase diagram of the TiNb system
is shown in Fig. 1.
Titanium alloys can be classified into three main categories
, + and -type, depending on alloy solute content and
heat treatment applied [4]. Beta-Ti alloys form one of the
most versatile groups of materials with respect to processing,

4868

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

Fig. 1. The phase diagram of the TiNb system.

microstructure and mechanical properties [4,7]. Titanium alloys


microstructure is a result of a number of solid/solid transformations. Firstly, the solidification process results in -phase
precipitation [1]. Beta grains develop during the cooling stage
along the solidification temperature range and the transus temperature. As a consequence, large sections, which cool slower,
have larger beta grains. The phase, typically plate-like shaped,
is formed along the grain boundaries when the sample is
cooled through the ( + ) phase field. Alpha plate colonies are
the transformation products of the phase when cooled below
the -transus temperature [1]. A number of studies have shown
that the resulting microstructure influences the final mechanical properties [1421]. In the case of Ti alloys, modifications
of microstructure by using several processes such as casting,
solution-treating, aging [1], quenching [7,15], roll bonding [16],
cold or hot working followed by heat treatment and forging [1]
have provided considerable changes in tensile strength, fracture
toughness, fatigue crack propagation, wear, corrosion and oxidation resistances and modulus of elasticity [13]. An increase
on Nb content tends to decrease the modulus of elasticity and
stabilizes the phase. Alloys with lower modulus of elasticity are nowadays desired in the search of similarity to that
of bone. Some studies have shown that TiNb alloys ranging
from 20 to 50 wt%Nb exhibit a modulus of elasticity of about
60 GPa [3,22,23], which is closer to that of bone when compared to those of other conventional Ti alloys, stainless steel
and CoCr alloys [14,7]. Afonso et al. [7] have recently analyzed the microstructures, phases and mechanical properties of
Ti20Nb alloy samples as a function cooling rate. They have
concluded that high cooling rates increase the Vickers hardness
and decrease the Young modulus (which is about 74 GPa for a
cooling rate of 160 K/s) [7]. They have also found a high volume
of martensite within the grains.
In terms of biomaterials applications, the most inconvenient
aspect is the degradation, which occurs due to the material

interaction with body or physiological fluids with concentration


of about 1 wt% NaCl [5,14,2426]. The corrosion resistance
can be considered a vital property for biomaterial components
and is associated with the problem of ion release of metallic
species which can be harmful for the organism. Although Ti
based alloys exhibit reasonable to good corrosion resistance due
to the titania formation on its surface, the nature, composition
and thickness of the protective oxide depend on the environmental conditions [27]. Generally, the degradation occurs on
biomaterial surfaces due to the chemical reactions between titania and chloride ions in the media. The degradation of TiNb
alloys results in a passive film which is mainly formed by Ti2 O3 , TiO2 (in rutile, brookite and/or anatase forms) and Nb2 O5
[25,28]. Tamilselvi et al. [29] have recently carried out corrosion
tests with Ti6Al4V and Ti6Al7Nb alloys in a saline solution
(0.9% NaCl) and have reported that the corrosion current density
decreases and the corrosion potential increases with increasing
impressed potentials, and that the latter alloy has a much better
corrosion resistance than the former one. Because of the concern about the biocompatibility of Al and V ions released from
metallic biomaterials, the development of new alloys without
such components in their composition is an essential task. Lopez
et al. [24] have carried out a comparative study on the corrosion behavior of Ti6Al7Nb, Ti13Nb13Zr and Ti4Nb15Zr
alloys in Hanks solution and they have reported very similar
current densities for all three Ti alloys experimentally examined.
The aim of this study was to investigate the effects of
two different heat treatments, i.e., furnace cooling and water
quenching, on the general electrochemical corrosion resistance
of Ti35 wt%Nb alloy samples immersed in a 0.9% NaCl
(0.15 mol L1 ) solution at 25 C and neutral pH range.
2. Experimental procedure
The Ti35 wt%Nb alloy samples were prepared from commercially pure metals: Ti (99.86 wt%) and Nb (99.99 wt%). A
non-consumable tungsten electrode furnace with a water-cooled
copper hearth under an argon atmosphere was used to generate the specimens. In order to homogenize the sample chemical
composition, at least five remelting steps were carried out and
in each operation the melt was slipped into the hearth. In all
the cases, an argon atmosphere and slight vacuum conditions
were employed. This procedure guarantees chemical homogenization of the solidification structure. After remelting steps, the
samples were solution-treated, i.e., heated at 1000 C (10 C)
for 24 h within the field, which is delimitated in the TiNb
phase diagram shown in Fig. 1, to create an homogeneous phase
and subjected to hot working by a rotary swaging process at
800 C (8 C). The samples were then machined to a diameter
of 12 mm (0.08), heat treated at 1000 C for 1 h and cooled
under two different cooling conditions: water quenched (WQ)
and furnace cooled (FC).
Specimens were longitudinally sectioned from the center of
the ingot samples, ground by using silicon carbide papers up
to 1200 mesh, polished and etched to reveal the microstructure by using Krolls etchant (5 mL of HF, 30 mL of HNO3 and
65 mL of H2 O). Microstructural characterization was performed

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874


Table 1
Chemical composition of NaCl (wt%)
Element
wt%

Ba
0.001

Br
0.01

K
0.005

Ca
0.005

PO4
5 ppm

Other
<2 ppm

by using optical microscopy (Olympus BX60M) and scanning


electron microscopy (SEM, Jeol JXA 840A) with energy dispersive X-ray detector (EDAX, NORAN, System Six 1.5, USA).
X-ray diffraction (XRD) measurements were carried out in order
to verify the phases constitution after the two different heat
treatments. X-ray diffraction patterns were obtained utilizing
a Rigaku diffractometer (DMAX 2200) operated at 40 kV and
30 mA with Cu K radiation and a wavelength, , of 0.15406 nm.
The working electrodes for the corrosion tests consisted of
TiNb alloy samples which were positioned at the glass corrosion cell kit, leaving a circular 1 cm2 metal surface in contact
with the electrolyte. The potential amplitude was set to 10 mV,
peak-to-peak (AC signal), with 5 points per decade and the
frequency range was set from 100 mHz to 100 kHz. The samples were further ground to a 1200 grit SiC finish, followed by
distilled water washing and air drying before measurements.
Electrochemical impedance spectroscopy (EIS) measurements began after an initial delay of 1 h for the sample to reach a
steady-state condition. The tests were carried out with the samples immersed in a solution of 0.9% NaCl (0.15 mol L1 ) under
a neutral pH range (of about 6.85), used to simulate the human
body fluid [5,14,2426], and under air atmosphere. The mean
chemical composition of NaCl used to prepare the 0.9 wt% NaCl
solution is shown in Table 1. A potentiostat (EG & G Princeton
Applied Research, model 273A) coupled to a frequency analyzer
system (Solartron model 1250), a glass corrosion cell kit with
a platinum counter-electrode and a saturated calomel reference
electrode (SCE) were used to perform the EIS tests.
Potentiodynamic measurements were also carried out in a
0.9% NaCl solution at 25 C using a potentiostat. These tests
were conducted by stepping the potential using a scanning rate of
0.2 mV/s from 0.800 mV (SCE) to +2800 mV (SCE). Using an
automatic data acquisition system, the potentiodynamic polarization curves were plotted and both corrosion rate and potential

4869

were estimated by Tafel plots by using both anodic and cathodic


branches. Duplicate tests for EIS and potentiodynamic polarization curves were carried out. In order to supply quantitative
support for discussions of these experimental EIS results, an
appropriate model (ZView version 2.1b) for equivalent circuit
quantification has also been used.
3. Results and discussion
3.1. Microstructures
Typical optical micrographs of furnace cooled (FC) and
water quenched (WQ) Ti35 wt%Nb alloy samples are shown in
Fig. 2(a) and (b), respectively. It can be seen that the FC sample
has a microstructure constituted by + phases. It is expected
the presence of the phase, an athermal-fragile phase [6], that
can improve some mechanical properties, such as the ultimate
and yield tensile strengths, Vickers hardness and modulus of
elasticity. On the other hand, the elongation can be significantly
decreased. Kim et al. have reported that silicon addition (of about
1 wt%) suppress the phase precipitation [4]. Moffat and Larbalestier [30] have reported the precipitation of phase due to
aging.
The WQ sample exhibits an acicular martensitic  phase disseminated in a matrix, as identified by the X-ray diffraction
patterns (XRD), shown in Fig. 3(a). Fig. 3(b) shows the EDAX
results confirming the average chemical composition of both
furnace cooled and water quenched samples. It is reported in the
literature that the amount of  martensitic phase increases with
increasing cooling rate due to the transformation of phase
into  [7,30]. It has also been reported that the WQ induces
a microstructure with a decreased hardness (due to the acicular martensite) and with a lower modulus of elasticity when
compared to that of the FC sample [3,6,7,22,23].
3.2. EIS measurements
In order to analyze the effects of the two different heat treatments, the furnace cooling and the water quenching, on the
general corrosion resistance of the Ti35 wt%Nb alloy samples,

Fig. 2. Typical Ti35 wt%Nb alloy sample in: (a) furnace cooled and (b) water quenched conditions (magnification 200; etchant: Krolls).

4870

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

Fig. 3. (a) X-ray diffraction (XRD) patterns and (b) EDAX results for furnace cooled and water quenched conditions.

EIS and polarization tests were performed. The results have


shown that both the modulus of impedance (Z) and the phase
angle (), which are represented as a function of frequency,
are lower for the quenched sample. Higher impedance and
phase angle are conducive to a nobler electrochemical behavior
[20,31]. After polarization significant displacements on the modulus of impedance (of about 3 106 ), for both furnace cooled
and quenched samples, have been clearly observed. It seems that
after polarization, the passive oxide film is thicker than after the
first EIS measurements. However, in order to predict the level
of oxide film protection and to quantify the experimental EIS

results, an equivalent circuit analysis has also been carried out.


According to the literature [5,3033], the film formed on Ti
alloys is composed by a bi-layered oxide consisting of a porous
outer layer and a barrier inner layer.
3.3. Equivalent circuit analysis
An appropriate model (ZView version 2.1b) for equivalent circuit quantification has been used. Fig. 4 shows the
experimental and simulated curves for both FC and the WQ
Ti35 wt%Nb alloy samples and the equivalent circuit chosen

Fig. 4. Typical experimental data and simulated curves by using the ZView software: (a) and (b) for the furnace cooled (FC) and (c) and (d) for the water quenched
(WQ) Ti35 wt%Nb alloy samples and (e) equivalent circuit for modeling impedance data performed after 1h in a 0.9% NaCl solution.

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

to fit the experimental data. The proposed equivalent circuit


shown in Fig. 4(e) is widely used for Ti alloys [5,10,11,30,33].
The agreement between experimental and simulated results
indicates that the experimental results are well fitted to the
proposed equivalent circuit. The fitting quality was evaluated by
chi-squared (2 ) [5,31,3439] values of about 104 , which were
interpreted by the ZView software and are shown in Table 2.
The physical significance of the circuit elements for the
proposed model has been previously reported in the literature
[5,10,3039]. The model assumes that the oxide layer on the
Ti35Nb alloy consists of a barrier-like inner layer and a porous
outer layer. In this model, Rel corresponds to the resistance
of the electrolyte (0.9% NaCl solution). R1 and R2 are the
resistances of the porous and barrier layers. For the mathematical analysis of impedance diagrams, a constant phase element,
CPE, was used instead of an ideal capacitor and ZCPE(1) ,
ZCPE(2) are defined as capacitances of porous and barrier layers, respectively. The parameter ZCPE generally denotes the
impedance of a phase element as ZCPE = [C(j)n ]1 (where
is frequency and 1 n 1) and corresponds to the capacitance (1 Hz < F < 1 kHz) [3039]. All impedance parameters are
shown in Table 2. These impedance parameters have confirmed
the qualitative analysis of EIS results.
The impedance parameters of TiNb alloy samples after 1 h
of immersion and after polarization in a 0.9% NaCl solution
indicate that the passive films which are formed are essentially
of the same nature. It seems that these oxide films correspond
to oxidation of TiO or Ti2 O3 to TiO2 (more stable). A compact
passive film provides a typical high capacitive behavior showing
a phase angle close to 90o , as observed in Fig. 4(b) and (d).
Fig. 5 shows typical SEM images after polarization in a 0.9%
NaCl solution and its correspondent EDAX patterns confirming
the presence of oxide formation for both FC (Fig. 5(a)) and
WQ (Fig. 5(b)) samples, respectively. It can also be observed a
slight decrease in both Ti and Nb contents compared to the mean
nominal chemical composition.
In the proposed model, the passive film consists of the inner
barrier (R2 ) and the porous (R1 ) layers. Comparisons between
capacitances (ZCPE(1) ) of porous and barrier layers (R1 and R2 ,
respectively) for FC and WQ samples after 1 h immersion in a
0.9% NaCl solution, permit to observe that the WQ sample has
a higher capacitance (of about 14 F cm2 ) and lower porous
and barrier resistances (R1 = 1.16  and R2 = 0.55 ) than those
of the FC sample, as shown in Table 2. This indicates that

4871

the WQ sample has poorer electrochemical corrosion behavior


[3036].
On the other hand, polarized FC and WQ samples have
impedance parameters which are very similar, permitting to conclude that similar corrosion resistances are expected. However,
as shown in Table 2, the capacitances, ZCPE(1) and ZCPE(2) , have
significantly decreased (of about 3 times) and the inner barrier layer resistances (R2 ) have increased (of about 3 times),
when comparing the results of samples after 1h of immersion
with that of the polarized samples. This indicates that the thicknesses of the oxide films or the oxide density of the inner
barrier layers have increased [39]. It seems that outer oxide
particles have migrated toward the inner layer reinforcing its
density. This has resulted in reduced R1 values for both polarized FC and WQ samples, as can be seen in Table 2. As a
direct consequence, a highly compacted passive Ti oxide film has
been formed. These impedances parameters permit to conclude
that both inner and outer layers can provide corrosion protection. This is slightly different from those impedance parameters
results obtained for TiAlNb, TiAlV and TiNbZr alloys
[5] which have shown that the corrosion protection is due to
the barrier layer. It is well known that the porous outer layer is
the responsible to osseointegration between implant and human
bone.
3.4. Potentiodynamic results
Fig. 6(a) and (b) shows experimental potentiodynamic polarization results, in a 0.9% NaCl solution, for the WQ and FC
samples after first and second polarization cycles, respectively,
and at the same position. This procedure was performed in order
to analyze the continuity, stability and intensity of the passive
Ti oxide film formation. The current densities (i) were obtained
from the polarization curves by Tafel plots using both cathodic
and anodic branches of the polarization curves. Such results reinforce the corrosion resistance tendency favoring the FC Ti35Nb
alloy sample which has been observed when analyzing both the
EIS diagrams and the equivalent circuit results. The experimental current density and potential obtained in this investigation, in
a 0.9% NaCl test solution at 25 C, are very similar to the experimental results obtained by Assis et al. [5] in a Hanks solution for
three different Ti alloys. In the present investigation, the current
densities (corrosion rates) for the FC sample are lower and the
potentials are nobler than those of the WQ sample after 1st and

Table 2
Impedance parameters for furnace cooled and water quenched Ti35 wt%Nb alloy samples after 1 h of immersion and polarized in a 0.9% NaCl test solution
Parameters

FC

WQ

FC (polarized)

WQ (polarized)

Rel ()
ZCPE(1) (F cm2 )
ZCPE(2) (F cm2 )
n1
n2
R1 () (porous)
R2 () (barrier)
2

279.5
10.98 (0.7)
7.48 (0.4)
0.92
0.84
2.30 106
1.43 106
16 104

279.8
14.11 (0.8)
6.46 (0.3)
0.88
0.74
1.16 106
0.55 106
11 104

276.2
3.70(0.3)
1.09 (0.1)
0.94
0.78
0.73 106
3.11 106
38 104

279.4
3.45 (0.3)
1.65 (0.1)
0.92
0.69
0.98 106
2.97 106
33 104

Values in parenthesis: error resulted from fitting the experimental data.

4872

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

Fig. 5. Typical SEM images after polarization in a 0.9% NaCl solution and EDAX patterns.

2nd polarization cycles, as shown in Fig. 6(a) and (b). For the
WC sample the potential is slightly displaced toward less noble
values and the current density is higher when compared to the
corresponding values of the FC sample, i.e., for the FC and WQ
samples the current densities and potentials are 0.044 A/cm2
and 363 mV; and 0.050 A/cm2 and 488 mV, respectively.
A partial stabilization on the current density is observed
at about 2 108 A cm2 for both potentials of FC and WQ

Ti35Nb alloy samples, which indicates that a Ti oxide film


has been formed [5]. However, this oxide film becomes more
stable at about 40 mV and 400 mV for WQ and FC samples,
respectively. These potentials characterize passive current densities (IPP ) of 8.30 A cm2 and of 8.05 A cm2 after the 1st
and the 2nd polarization cycles, respectively. Although these
IPP currents show a slight linear increasing tendency, significant
breakdowns in the passive film have not been verified. Com-

Fig. 6. Potentiodynamic polarization curves of furnace cooled (FC) and water quenched (WQ) Ti35 wt%Nb alloy samples after: (a) first and (b) second polarization
cycle in a 0.9% NaCl solution.

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

parisons between the potentiodynamic polarization curves for


FC and WQ samples show that the FC sample exhibits a lower
current density (of about 1530% lower) associated to a potential nobler than those of the WQ sample in both 1st and 2nd
polarization cycles, as shown in Fig. 6. After the 2nd polarization cycle (Fig. 6(b)) at about 2 108 A cm2 , the potentials
for both FC and WQ samples rapidly increase when compared
to the corresponding potentials after the 1st polarization cycle
(Fig. 6(a)). A possible replacement from a less stable to a more
stable condition or nobler protective passive film occurs at potentials between 0 and 40 and 0400 mV for the WQ and the FC
samples, respectively.
The passive current densities (IPP ) for both FC and WQ
Ti35 wt%Nb alloy samples (of about 8 A cm2 ) are higher
than those reported by Assis et al. in a recent study [5] of commercial Ti13Nb13Zr, Ti6Al4 V and Ti6Al7Nb alloys and
are similar to the IPP values obtained by Martins et al. [39]
in another recent study with a Ti30Nb7.5Zr alloy (of about
7 A cm2 ). Although a very protective Ti oxide film is formed
on the surface of such samples, the high values of IPP suggest
that this oxide can be more defective or porous, which can be
helpful to the osseintegration process.
3.5. Resulting microstructure and electrochemical
corrosion behavior
Experimental results of EIS and polarization measurements
associated to an equivalent circuit analysis have shown that
the operational conditions of heat treatment can be used as an
alternative way to control the resulting microstructure which
affects the electrochemical corrosion behavior of Ti alloys. In
the present investigation, the results of corrosion tests have indicated that quenching of a Ti35 wt%Nb alloy sample decreases
the electrochemical performance when compared to that of the
furnace cooled (FC) sample. Water quenching Ti-based alloys
results in a microstructure formed by an acicular martensitic
 -phase (orthorhombic) in a matrix [15], while the FC
induces a microstructure consisting of ( + ) phases and a
slight amount of -phase, as aforementioned and shown in
Fig. 2. The  -phase is constituted by fine needle-like traces
of martensite which is associated to a residual stress-induced
martensitic transformation. Although the phase (the nobler
matrix phase) envelopes the  phase (the less noble acicular
phase) avoiding or minimizing the galvanic corrosion effect, the
stress-induced martensitic transformation displaces the corrosion potential to more active behavior and increases the corrosion
rate. On the other hand, the FC sample has a microstructure
formed by a more homogeneous , and phases distributed
in larger grains (with mean diameters >200 m) without the
occurrence of the afore-mentioned deleterious effects of stressinduced martensitic transformation and micro cells galvanic
corrosion.
4. Conclusions
The following main conclusions can be drawn from the
present experimental investigation:

4873

(1) Both water quenching (WQ) and furnace cooling (FC) heat
treatments of Ti35 wt%Nb alloy samples have provided a
passive film formation during corrosion tests in a 0.9% NaCl
solution, which has been observed by the very low experimental current densities. After a second polarization cycle,
the corrosion potentials for both (FC) and (WQ) samples
were displaced toward less noble conditions while the corrosion rates have decreased due to a thicker Ti oxide film
formation.
(2) The experimental EIS diagrams, potentiodynamic polarization curves and the equivalent circuit parameters have shown
that the WQ Ti35Nb alloy sample had the electrochemical
corrosion resistance decreased compared to that of the FC
sample due to stress-induced martensitic transformation.
(3) The EIS results and the equivalent circuit analysis have indicate that the protective passive films formed on both FC and
WQ Ti35Nb alloy samples are composed of a dual-layered
oxide consisting of an inner barrier layer associated to high
impedance and responsible for corrosion protection, and an
outer porous layer, of lower impedance, which apparently
facilitates the osseointegration.
Acknowledgements
The authors acknowledge financial support provided by
FAPESP (The Scientific Research Foundation of the State of Sao
Paulo, Brazil), FAEPEX-UNICAMP and CNPq (The Brazilian
Research Council).
References
[1] J.R. Newman, D. Eylon, J.K. Thorne, Titanium and Titanium Alloys, Metals Handbook, ASM International, Metals Park, OH, USA, 1988, p. 824.
[2] C. Eriksson, K. Ohlson, K. Richter, N. Billerdahl, M. Johansson, H. Nygren,
J. Biomed. Mater. Res. 83A (2007) 1062.
[3] E.B. Taddei, V.A.R. Henriques, C.R.M. Silva, C.A.A. Cairo, Mater. Sci.
Eng. 24C (2004) 683.
[4] H.S. Kim, S.H. Lim, I.D. Yeo, W.Y. Kim, Mater. Sci. Eng. 449451A (2007)
322.
[5] S.L. Assis, S. Wolynec, I. Costa, Electrochim. Acta 51 (2006) 1815.
[6] Y. Mantani, M. Tajima, Mater. Sci. Eng. 442A (2006) 409.
[7] C.R.M. Afonso, G.T. Aleixo, A.J. Ramirez, R. Caram, Mater. Sci. Eng.
27C (2007) 908913.
[8] T. Akahori, M. Niinomi, Mater. Sci. Eng. A 243 (1998) 237.
[9] R.J. S Solar, Corrosion resistance of titanium surgical implant alloys: a
review, in: Corrosion and Degradation of Implant Materials STP 684,
American Society for Testing of Materials, 1979, p. 259.
[10] R. Karpagavalli, A. Zhou, P. Chellamuthu, K. Nguyen, J. Biomed. Mater.
Res. 83A (2007) 1087.
[11] N. Padilla, A. Bronson, J. Biomed. Mater. Res. 81A (2007) 531.
[12] M.A. Khan, R.L. Williams, D.F. Williams, Biomaterials 20 (1999) 631.
[13] L.J. Xu, Y.Y. Chen, Z.G. Liu, F.T. Kong, J. Alloys Compd. 453 (2006)
320324.
[14] A. Choubey, R. Balasubramaniam, B. Basu, J. Alloys Compd. 381 (2004)
288.
[15] Y. Mantani, M. Tajima, Mater. Sci. Eng 43440A (2006) 315.
[16] R.G. Zhang, V.L. Acoff, Mater. Sci. Eng. 463A (2007) 67.
[17] N.J. Petch, J. Iron Steel Inst. 174 (1953) 25.
[18] P. Donelan, Mater. Sci. Technol. 16 (2000) 261269.
[19] P.R. Goulart, J.E. Spinelli, W.R. Osorio, A. Garcia, Mater. Sci. Eng. 421A
(2006) 245.

4874

A. Cremasco et al. / Electrochimica Acta 53 (2008) 48674874

[20] W.R. Osorio, P.R. Goulart, G.A. Santos, C. Moura Neto, A. Garcia, Metall.
Mater. Trans. 37A (2006) 2525.
[21] C.J. Boehlert, C.J. Cowen, J.P. Quast, T. Akahori, M. Niinomi, Mater. Sci.
Eng. 28 (2008) 323330.
[22] Y.H. Hon, J.Y. Wang, Y.N. Pan, Metall. Mater. Trans. 44 (2003) 2384.
[23] C.M. Lee, C.P. Ju, J.H.C. Lin, J. Oral Rehabil. 29 (2003) 314.
[24] M.F. Lopez, A. Gutierrez, J.A. Jimenez, Electrochim. Acta 47 (2002) 1359.
[25] Y.F. Zheng, B.L. Wang, J.G. Wang, C. Li, L.C. Zhao, Mater. Sci. Eng. 438A
(2006) 891.
[26] R.M. Bezerra, P.C.R.D. Souza, I. Ramires, M.A. Bottino, A.C. Guastaldi,
Ecl. Quim. 24 (1999) 113.
[27] I. Gurrapa, Mater. Charact. 51 (2003) 131.
[28] M. Anuwar, R. Jayaganthan, V.K. Tewari, N. Arivazhagan, Mater. Lett. 61
(2007) 1483.
[29] S. Tamilselvi, R. Murugaraj, N. Rajendran, Mater. Corrosion 58 (2007)
113.

[30] D.L. Moffat, D.C. Larbalestier, Metall. Trans. 19A (1988) 1687.
[31] W.R. Osorio, J.E. Spinelli, I.L. Ferreira, A. Garcia, Electrochim. Acta 52
(2007) 3265.
[32] J. Pan, D. Thierry, C. Leygraf, Electrochim. Acta 41 (1996) 1143.
[33] S. Gudic, J. Radosevic, M. Kliskic, Electrochim. Acta 47 (2002) 3009.
[34] W.R. Osorio, J.E. Spinelli, N. Cheung, A. Garcia, Mater. Sci. Eng. 420A
(2006) 179.
[35] W.R. Osorio, N. Cheung, J.E. Spinelli, P.R. Goulart, A. Garcia, J. Solid
State Electrochem. 11 (2007) 1421.
[36] W.R. Osorio, L.R. Garcia, P.R. Goulart, A. Garcia, Mater. Chem. Phys. 106
(2007) 343.
[37] W.R. Osorio, P.R. Goulart, A. Garcia, Mater. Lett. 62 (2008) 365.
[38] W.R. Osorio, C.M.A. Freire, A. Garcia, J. Alloys Compd. 379 (2005)
179.
[39] D.Q. Martins, W.R. Osorio, M.E.P. Souza, R. Caram, A. Garcia, Electrochim. Acta 53 (2008) 2809.

Potrebbero piacerti anche