Você está na página 1de 27

Author's Accepted Manuscript

Influence of phase transformations on


dynamical elastic modulus and anelasticity of beta Ti-nb-Fe alloys for biomedical
applications
J.M. Chaves, O. Florncio, P.S. Silva Jr., P.W.
B. Marques, C.R.M. Afonso

www.elsevier.com/locate/jmbbm

PII:
DOI:
Reference:

S1751-6161(15)00073-9
http://dx.doi.org/10.1016/j.jmbbm.2015.02.030
JMBBM1407

To appear in: Journal of the Mechanical Behavior of Biomedical Materials

Received date:5 December 2014


Revised date: 21 February 2015
Accepted date:
26 February 2015
Cite this article as: J.M. Chaves, O. Florncio, P.S. Silva Jr., P.W.B. Marques,
C.R.M. Afonso, Influence of phase transformations on dynamical elastic
modulus and anelasticity of beta Ti-nb-Fe alloys for biomedical applications, Journal of the Mechanical Behavior of Biomedical Materials, http://dx.doi.org/
10.1016/j.jmbbm.2015.02.030
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early
version of the manuscript. The manuscript will undergo copyediting,
typesetting, and review of the resulting galley proof before it is published in
its final citable form. Please note that during the production process errors
may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

Influence of phase transformations on dynamical elastic modulus and


anelasticity of beta Ti-Nb-Fe alloys for biomedical applications
J. M. Chavesa*, O. Florncioa, P. S. Silva Jr.a, P. W. B. Marquesa, C. R. M. Afonsob
a

Department of Physics, UFSCar, C.P. 676, CEP 13565-905, So Carlos-SP, Brazil.

Department of Materials Engineering, DEMa-UFSCar, CEP 13565-905, So Carlos-SP, Brazil.

Abstract
Recent studies in materials for biomedical applications have focused on -titanium alloys that
are highly biocompatible, free of toxic elements and with an elastic modulus close to that of
human bone (10-40 GPa). Beta Ti-xNb-3Fe (x = 10, 15, 20 and 25 wt.%) alloys were obtained
by rapid solidification and characterized by anelastic relaxation measurements at temperatures
between 140 K and 770 K, using a free-decay elastometer, as well as analysis by Differential
Scanning Calorimetry (DSC), X-ray Diffraction (XRD) and Scanning Electron Microscopy
(SEM) and Transmission Electron Microscopy (TEM). The observed stabilization of the -phase
with rising Nb content was linked to the strength of the relaxation peak around 570 K. The phase
transformations detected in the anelastic relaxation spectra agreed with those observed in the
DSC curves. However, the results from anelastic relaxation spectra provide more detailed
information about the kinetics of phase transformations. At temperatures between 140 K and 300
K, there was an indication of a reversible transformation in the alloys studied. The elastic
modulus measurements showed a hardening of the material, between 400 K and 620 K, related to
the -phase precipitation. However, the starting temperature of -phase precipitation was clearly
influenced by the Nb content, showing a shift to high temperature with increasing percentage of
Nb. At temperatures above 620 K, a fall was observed in the dynamical elastic modulus,
accompanied by a relaxation peak centered at 660 K, which was attributed to the growing phase arising from the -phase, which acts as a nucleation sites or from the decomposition of the
metastable -phase. XRD patterns confirmed the formation of , and phases after mechanical
relaxation measurements. A predominant phase with dendritic morphology was observed,
which became more stable with 25 wt.% Nb. The lowest elastic modulus was of 65 GPa obtained
in the Ti-25Nb-3Fe alloy, representing a good low value for a -Ti alloy with a relatively low
addition of stabilizing elements (Nb and Fe).
Keywords: titanium alloys, anelastic relaxation, elastic modulus, phase transformation.
* Corresponding author: e-mail address: javier@df.ufscar.br

1. Introduction
Research on titanium alloys designed for biomedical applications has intensified a great deal in
recent decades, on account of their interesting combination of properties, such as excellent
chemical biocompatibility, low elastic modulus and corrosion resistance. These properties can be
tuned by the addition of - and -stabilizing alloying elements, adjusting the processing
parameters or heating and thermo-mechanical treatments (Niinomi, 2002). This tuning depends
on the close relationship between the microstructure and phases formed, their distribution, size
and morphology(Elmay et al., 2013). Thus, -Ti alloys free of toxic elements have been designed
in recent research, with low values of elastic modulus, which are closer to that of human bone
(10-40 GPa) that other metal alloys used as biomaterials: Co-Cr (220 GPa), stainless steels (190
GPa), CP-Ti (100 GPa) and + type Ti-6Al-4V (110 GPa) (Abdel-Hady Gepreel and Niinomi,
2013; Long and Rack, 1998; Niinomi, 2008). The similarity in the elastic modulus between
human bone and the implant is very important in alloys used as orthopedic implants, because a
great difference can lead to the stress-shielding phenomenon. This phenomenon, which relieves
the bone of external stress, reduces the reposition of the bone surrounding the metallic implant
and consequently leads to loss of the implant or new fracture of the bone after the implant is
removed (Niinomi and Nakai, 2011).
However, control of the crystalline phases is of great importance in the design of new
biomaterials alloys, which should favor the presence of metastable phases such as and
martensites, the latter responsible for the shape memory effect and superelasticity (Baker C,
1971; Kim et al., 2006); on the other hand, the -phase is deleterious to mechanical properties of
the -Ti alloys, increasing their hardness and elastic modulus and can occur during rapid cooling
(quenching), aging heat treatments or upon deformation (Lee et al., 2002b; Li et al., 2013;
Mantani and Tajima, 2006).
In this context, the mechanical spectroscopy technique can be very useful for the characterization
of dynamical processes, including phase transformations in -Ti alloys. Thus, internal friction is
a physical property very sensitive to processes that involve absorption of mechanical energy by
atomic rearrangement, phase transformations, matrix-solute interactions (interstitial and/or
substitutional), diffusion process and so on (Nowick and Berry, 1972; Schaller et al., 2001).
Ti-Nb alloys have been widely studied, on account of their low elastic modulus, shape memory
and superelasticity (Kent et al., 2013; Lopes et al., 2011; Tobe et al., 2013). It has been observed
that the mechanical properties, structure and morphology are sensitive to Nb content. In this
system, alloys containing around 15 wt.% or less of Nb exhibit martensite (hexagonal); at 17

wt.% - 25 wt.% Nb, the martensite phase (orthorhombic) predominates and, in alloys with
more than 30 wt.% Nb, the phase is dominant with equiaxal grain structure (Lee et al., 2002b).
Therefore, the study of the new or modified alloys by this technique can help to improve the
performance of the alloys used in biomedical applications.
The present study is focuses on the Ti-Nb-Fe system, in which the fractions of -stabilizing
elements Nb and Fe influence phase stability. Fe is added on the basis its strongly effect
stabilizing on the -phase, mechanical strengthening potential, low cost and its lowering effect
on the melting point of the alloy, which can facilitate the production of alloys, with the aim of
achieving a high performance-to-cost ratio. On the other hand, in compositions close to the
eutectic one, this alloy can provide a greater thermal supercooling from the liquid, favoring the
formation of refined microstructures, dendritic growth or metastable phases (Reed-Hill, 1973).
Thus, the anelastic behavior of the Ti-xNb-3Fe alloys (x=10, 15, 20 and 25 wt.%) was
characterized by mechanical spectroscopy and its correlation with the sequence of phase
transformations during cyclic heat treatment will be demonstrated, in order to map the
optimization of heat treatments for these -Ti alloys, so as to guarantee the desired phases and
microstructure, with the aim of biomedical application.
2. Materials and methods
Ti-xNb-3Fe alloys (x=10, 15, 20 and 25 wt.%) were processed from pure elements such as Ti
(sponge: 99.5%), Nb (99.8%) and Fe (99.97%). Ingots were fabricated in an arc furnace (AM
Arc Melter, Edmund Bhler) in an inert atmosphere of argon in a water-cooled copper hearth
crucible and then remelted to ensure compositional homogeneity. From these ingots, smaller
samples were rapidly solidified in a commercial arc furnace (EDG Discovery Plasma) by
suction casting-technique, the suction arising from a vacuum between the copper crucible (upper
chamber) and the copper mold (lower chamber), with shallow grooves for the vacuum suction.
The cooling rates imposed during the copper-mold casting are estimated at 101 K/s to 103 K/s.
The copper mold used plate shape, with varying thickness, had dimensions 20x50 mm2 and
thickness 2.0 mm, 1.0 mm and 0.5 mm. Plates were cut in dimensions of around 20x5x0.5 mm3
as specimens for mechanical spectroscopy tests.
Anelastic spectra (internal friction and elastic modulus relative variation to the room temperature
value (E/E) plotted against temperature) were obtained with a free decay elastometer
equipment (Vibran Technologies AE-102) from the logarithmic decay of free oscillations of the
first tone in flexural vibration mode. Dynamical elastic modulus (E) was calculated from the

resonance frequency of the fundamental mode of flexural vibration (f1), in the clamp-free
configuration, by the relationship (Nowick and Berry, 1972):
f1 = 0.1615

h
l2

(1)

where h is the thickness, l the length and the density of the sample.
Analysis in the free decay elastometer consisted of a series of cyclic measurements, during
which the sample was cooled, by immersing the chamber containing it in liquid nitrogen, from
room temperature (RT) down to ~140 K, then heated from 140 K up to 770 K, at a cooling and
heating rate of 1 K/min. In this equipment, the sample was held in a vacuum chamber with an
internal resistive furnace, at a pressure below 4x105 Torr. The data for the anelastic spectra were
collected in measurements made at each 0.5 K.
Complementary data on the crystalline structure in the as-quenched condition and after heat
treatments carried out during the anelastic relaxation measurements, were provided by X-ray
Diffraction (XRD) in a Siemens D5005 X-ray Diffractometer. The microstructure was
characterizated by Scanning Electron Microscopy (SEM) whit a JEOL JSM-5800LV and
Transmission electron microscopy (TEM) using a FEI Tecnai G2 F20 equipment with the 200
KV beam for field emission (FEG) coupled with energy dispersive spectroscopy (EDS) EDAX.
The samples were prepared metallographically by gridding, polishing and etching with Kroll
reagent (6 mL HNO3, 3 mL HF and 91 mL of H2O). Measurements of interstitial elements
content were carried out using a LECO TC-436 analyzer. Additionally, Vickers hardness
measurements were carried out with a Stiefelmayer KL2 microhardenss tester, taking the mean
value for six indentations with load of 200 gf during 15 s. To compare the kinetics of phase
transformations, analysis by Differential Scanning Calorimetry (DSC) was carried out in Netzch
STA 404 calorimeter during two thermal cycles between 140 K and 770 K at a heating rate of 1
K/min.

3. Results and discussion


3.1. Microstructural characterization and phase identification
Fig. 1 displays the micrographs of rapidly solidified Ti-xNb-3Fe (x=10, 15, 20 and 25 wt.%)
alloys, exhibiting an isotropic dendritic growth morphology, without preferred orientation, in
which the microstructure become more refined with increasing of Nb content. This refinement is
due to the effect of the -stabilizer eutectoid alloying element (Fe) that increases the thermal
supercooling that occur during rapid solidification.

The chemical analysis by EDS and the interstitials solute atoms content measured by gas
analysis, shown in tables 1 and 2, respectively, indicate that the preparation of the alloy was
adequate in regards as nominal composition.
Fig. 2 shows the XRD patterns of rapidly solidified Ti-xNb-3Fe (x=10, 15, 20 and 25 wt.%)
alloys in the as-quenched condition. In the diffractograms there are typical diffraction peaks of
the -phase (bcc) stabilized at room temperature, besides diffraction peaks of the metastable phase (hcp) for the Ti-10Nb-3Fe and Ti-15Nb-3Fe alloys. It can be seen that increasing the Nb
percentage from 10 to 25 wt.% stabilized the -phase formed. Additionally, the presence of 3
wt.% Fe, a strong -stabilizing eutectoid element, favored the stabilization of the -phase at just
20 wt.% Nb, instead of a minimum of 35 wt.% Nb to obtain the stable -phase in the binary TiNb system (Lee et al., 2002a; Lin et al., 2002).
Regarding the splitting in -Ti phase peak, observed in the Fig. 2, is due to segregation in
dendritic growth, according to SEM images in BSE (backscattered electrons) mode (Fig. 3),
resulting in variation of crystalline cell parameter with Nb-rich -Ti phase (brighter) in the inner
region of the dendrites and Ti and Fe-rich -Ti phase (darker) in the interdendritic region of the
microstructure. That differ from -Ti phase separation that occurs in stable -Ti alloys
(nanometric scale) with higher contents of -stabilizing elements, such as TNZT, depending on
the processing route and heat treatment conditions (Afonso et al., 2010).
Thermal analysis by DSC, collected at a heating rate of 1 K/min, of samples with 10 and 15
wt.% Nb (Fig. 4) showed an exothermic process around 790 K, whereas during the second
heating cycle, no process was revealed. For samples with 20 and 25 wt.% Nb, an endothermic
process around 715 K was observed, and in the second heating cycle the process still takes place.
From literature data (Mantani and Tajima, 2006), precipitation of the -phase is associated with
the exothermic peak and precipitation of and -phases with the endothermic peak.
3.2. Anelasticity and dynamic elastic modulus characterization
In Fig. 5, the elastic modulus is plotted against electron concentration (e/a), it can be observed
that the addition of Nb leads to an increase in the (e/a) ratio, which is linked to a decrease in the
elastic modulus. This fact is related to -phase stabilization, since it has been suggested in
(Ikehata et al., 2004) that, for e/a values less than 4.20, the bcc structure is unstable and, on
decreasing the valence electron concentration, the stability of the hcp structure rises, so that
metastable phases, such as or martensites, can be formed, increasing the elastic modulus. For

this alloy system, higher stabilization of the -phase was achieved from e/a= 4.23 (20 wt.% Nb)
upwards, while at lower values the -phase appears, as shown in XRD patterns. This result is
consistent with that established by Collings (Boyer R. et al., 1994), who showed that in Ti-TM
alloys (TM= transition metal), the -phase forms when the content of the alloy lead to an
electron concentration (e/a) between 4.12 and 4.21.
Comparing these results with those obtained by Fedotov (Collings, 1986) in the Ti-Nb system,
and the curve Ti-TM for binaries systems proposed by Collings (Collings, 1984), it can be seen
that the points for the Ti-Nb-Fe alloy system are to the left of the curves for binary systems,
indicating the higher stability of the -phase, caused by suppression of the metastable phases
and . Thus, the 3 wt.% Fe leads to a stabilization of the -phase in samples with low Nb
content, a fact linked to the higher e/a ratio in the Ti-Nb-Fe system.
However, TEM micrographs of Ti-10Nb-3Fe and Ti-25Nb-3Fe samples in the as-quenched
condition, in Figs. 6 and 7 respectively, showed a nanoscale structure. Fig. 6 shows a bright field
(BF) image (Fig. 6.a) and dark field (DF) image (Fig. 6.b) of athermal nanometric -phase
precipitate dispersed in -Ti (bcc) grain matrix, due to rapid quenching from the melt. The
respective selected area electron diffraction (SAED) pattern of Ti-10Nb-3Fe (inset) is shown
with orientation relationship of zone axis [2 1 2]//[001]. Fig. 7 shows TEM micrographs of the
Ti-25Nb-3Fe sample, with BF (Fig. 7.a) and DF micrographs (Fig. 7.b) showing athermal phase precipitates, together with respective SAED pattern (inset Fig.7.a) with orientation
relationship of zone axis [1 1 5]. Although XRD patterns of a Ti-25Nb-3Fe sample in Fig. 2 do
not have clear peaks of -phase, TEM analysis confirms that there is still some fraction of this
phase precipitated on the nanoscale. It can be seen as well that the -phase is finer in the Ti10Nb-3Fe alloy and in a greater fraction than that in the Ti-25Nb-3Fe alloy, owing to Nb
content. Thus, the elastic modulus values and e/a ratio are consistent with the -phase fraction in
alloys Ti-10Nb-3Fe and Ti-25Nb-3Fe.
3.2.1. Anelastic relaxation spectra for Ti-10Nb-3Fe alloy
Fig. 8 shows the anelastic relaxation spectra (elastic modulus relative variation to the room
temperature value (E/E) and internal friction, plotted against temperature) and XRD patterns
for the Ti-10Nb-3Fe alloy during three measurement cycles between 140 K and 770 K.
In these spectra a relaxation structure can be seen that depends on the temperature and is related
to phase transformations, since heat treatment of metastable phases promotes changes in the

system toward an energetically more favorable equilibrium. These spectra show more detailed
evidence of the kinetic of phase transformations than the DSC curves.
In the first thermal cycle at low temperature (300 K to 140 K), there is a relaxation structure that
represent high absorption of elastic energy. In this step, the internal friction (Q-1) reaches values
around 4x10-3, and that energy is recovered in the subsequent heating up to room temperature.
Similar behavior is observed in the elastic modulus, as its variation does not exceed 1%. This
indicates that the process is almost completely reversible, so that it must be related to the reverse
martensite transformation of the phase, in agreement with the literature (Bertrand et al.,
2013); the reversible movement of the lattice by austenite/martensite transformation dissipates
mechanical energy constituting a source of structural damping. These observations indicate that
this alloy probably display a shape memory effect in this range temperature, but more specific
study is necessary and this will be discussed in a forthcoming publication. At higher
temperatures, a minimum in E/E is reached around 400 K; this behavior may be associated with
some characteristic temperature of this phase transformation.
In the temperature range between 400 K and 700 K, during the first thermal cycle, E/E reaches
a maximum value of around 10% at 615 K that could be related to the precipitation or growth of
the metastable -phase that provokes the hardening of the alloy, since, as reported frequently in
the literature (Li et al., 2013; Tane et al., 2013), this phase has the highest elastic modulus among
the phases formed in -Ti alloys. At higher temperatures E/E decreases, while Q-1 continues to
rise; this behavior is associated with the growth of -phase. In the subsequent cooling, a
hardening of the alloy is observed, achieving 20% of variation on E/E, as can be observed from
the beginning of the second thermal cycle at room temperature. The phase transformations are
confirmed by XRD patterns after first thermal cycle, as observed in Fig 8d.
In the following heat treatment, it is observed that the presence of and phases affect
considerably the reversible behavior at low temperature, since Q-1 decreases to 1x10-3 and
0.5x10-3 during the second and third thermal cycle, respectively. The final E/E value, after third
thermal cycle, was around 12%, corresponding to 84 GPa.
The intermediate value for E/E could be associated with competition among the series of
phases transformations, in sequence: , leading from the initial microstructure + to
a final microstructure composed of +. This transformation sequence has already been reported
in the literature (Chaves et al., 2014; Lopes et al., 2011; Ohmori et al., 2001), since that
metastable -phase (isothermal) can precipitate under aging heat treatments or during cooling at

lower cooling rates from the -phase field. Besides that, growth of -phase can occur from phase that acts as a nucleation site, or from the decomposition of the -phase metastable.

3.2.2. Anelastic relaxation spectra for Ti-15Nb-3Fe alloy


Anelastic relaxation spectra E/E and Q-1 plotted against temperature) and XRD patterns for the
alloy Ti-15Nb-3Fe, during three measurement cycles between 140 K and 770 K, are shown in
Fig. 9.
The alloy Ti-15Nb-3Fe behaves similarly at low temperature, to Ti-10Nb-3Fe. In the first
thermal cycle, a reversible process occurs and it is affected by the appearance of new phases, as
well as changes in the existing phases in each new cycle. Also the weak minimum in E/E is
shifted to around 425 K during first heating, which may be due to the higher proportion of phase stabilized.
In this alloy, at temperatures above 425 K, the hardening due to the growth of nanometric phase is not clearly observed, as E/E reaches an apparent maximum of 5% around 580 K and
then decreases at higher temperatures. This confirms that Ti-15Nb-3Fe has a -phase and a lower
proportion of -phase, leading to a rapid transformation and thus no apparent rise in of the

-phase fraction during heating cyclic. Thus, -phase is formed by the decomposition of
metastable -phase. In this way, the initial + microstructure is transformed to the equilibrium

+ microstructure during the thermal cycles in the anelastic relaxation measurements. The
addition of Fe as a strong eutectoid stabilizer of -Ti acted as a strong suppressor of precipitation
of the -phase, even at low content (3 wt.%.), preventing the massive -phase formation. The
phases transformations were confirmed by XRD patterns observed in Fig. 9.d. Thus, the elastic
modulus remained almost constant, reaching a final value of 67 GPa.
3.2.3. Anelastic relaxation spectra for Ti-20Nb-3Fe alloy.
For Ti-20Nb-3Fe alloy, in the anelastic relaxation spectra in Fig. 10, the maximal relaxation
strength at the minimum temperature was less intense, since the Q-1 maximum value at 140 K
was 3x10-3 and little affected by the subsequent thermal cycles. At temperatures between 400 K
and 550 K, a slight hardening of material is observed in E/E and, above this temperature, there
is a sharp increase in E/E that reaches a maximum of 18% at 630 K. As stated above, this
behavior arises from the transformation, which can be favored by the low heating rate (1
K/min) during the thermal cycle. In the subsequent thermal cycles, a hardening of up to 25% was
observed, which stabilized at 20% in the final cycle. This behavior may be associated firstly with

the transformation of metastable -phase, from which the -phase nucleated and, after the phase is formed and nucleated, as it was reported that, in -Ti alloys, -phase laths are nucleated
at the / interfaces and grow into both the matrix and particles, thus consuming the -phase
particles(Ohmori et al., 2001). The evolution of phases in the sample led the original -phase to a
structure +, as shown in XRD patterns in each thermal cycle (Fig. 10d).
Internal friction (Q-1) curve for the first thermal cycle, in the range 400 K to 750 K shows two
anelastic relaxation peaks of strength 2x10-3, centered at 570 K and 680 K, respectively. In the
subsequent cooling, the curves does not replicate this behavior, as only the peak at higher
temperature appears. In the second and third thermal cycles the peak is centered around 635 K
and the relaxation strength falls to 1.6x10-3 and 1.4x10-3, respectively. In addition, a weak
inflection is seen in E/E, at the maximum strength of the relaxation; this behavior is
characteristic of Snoek-type relaxation, according to the standard anelastic solid theory (Nowick
and Berry, 1972). The nature of the relaxation peak will be discussed in section 3.3.
3.2.4. Anelastic relaxation spectra for Ti-25Nb-3Fe alloy
The behavior of the anelastic spectra for Ti-25Nb-3Fe alloy submitted to three thermal cycles is
presented in the Fig. 11.
The lower temperature behavior seen in the previous alloys of system Ti-xNb-3Fe (x = 10, 15,
20 wt.%) is again similar for this composition, where the relaxation strength in Q-1 is affected by
each thermal cycle. In the first cycle, the E/E curve shows a weak fall (4%) that stabilizes
between 425 K and 550 K, above which E/E rises sharply, reaching a hardening of 20% at 640
K, followed by a drop up to 750 K. In the subsequent cooling E/E increased slowly, reaching
10%. In the second thermal cycle, contrary to what happened at other compositions, around 650
K a new maximum of 8% is formed in E/E. This is possibly related to residual -phase in the
specimen. In the third cycle, E/E stabilizes around 23%, corresponding to 79 GPa .
In the Q-1, during the first cycle, a distinctive relaxation anelastic peak is displayed around 565
K, with a relaxation strength of 5.5 x10-3, much higher than in Ti-20Nb-3Fe alloy. At higher
temperatures, a shoulder appears at 680 K, similar to the second peak for Ti-20Nb-3Fe. In the
subsequent thermal cycles, the relaxation peaks overlap to form a single peak with relaxation
strength reduced to 2.5x10-3 and centered around 620 K. This behavior will be discussed in the
next section.

In table 3 are summarized the elastic modulus and microhardness values, as well as the phases
observed in alloys Ti-xNb-3Fe (x= 10, 15, 20, 25 wt.%) in each thermal cycle. Thus, it is seen
that the -phase is responsible for the higher values of microhardness and elastic modulus.
3.3. Relation of anelastic relaxation peak with structure phases
In Fig. 12, it can be seen that the strength of the relaxation peak around 565 K depends on the Nb
content in the system Ti-xNb-3Fe (x= 10, 15, 20, 25 wt.%).
In section 3.2 (Fig. 5), it was observed that the addition of Nb leads to a decrease in the elastic
modulus, since the -phase is more stable. Therefore, the height of the relaxation peak may be
linked to the stability and amount of -phase, related to the Nb content in the system. Thus, it can
be observed that in the alloys containing 10 and 15 wt.% Nb, which show + phases in the
structure, the peak height is imperceptible, compared to alloys containing 20 and 25 wt.% Nb,
which have a small amount of -phase, as shown by the XRD and TEM results (Figs. 2 and 7
respectively). Similar behavior was observed by Zhou and coworkers in Ti-Nb alloys (Zhou et
al., 2011).
The second peak, around 680 K, may be associated with the growth of -phase, since this peak
appears in the region where E/E drops sharply, so that the -phase may be emerge from the phase, which acts as a nucleation site or by decomposition of the metastable -phase(Ohmori et
al., 2001). Curiously, this drop in E/E for Ti-15Nb-3Fe alloy begins at a lower temperature,
which suggests that the -phase arises without hardening from the -phase.
In the E/E curve (Fig. 12), it is observed that the onset temperature for hardening shifts to
higher temperatures with increasing Nb content. This behavior indicates that, with 3 wt.% Fe in
the alloys, the addition of Nb, retards the growth of -phase, which develops rapidly when it
reaches a characteristic temperature, for example 550 K in the alloys containing 20 and 25 wt.%
Nb.
In the subsequent cycles, for alloys containing 20 and 25 wt.% Nb, relaxation peak overlap, their
intensity decreases and an inflection is seen in E/E, characterizing a Snoek-type relaxation,
mentioned above. Other authors (Almeida et al., 2009; Lu et al., 2012; Yin et al., 2006) have
reported a Snoek-type relaxation in -Ti alloys, suggesting that this peak can be caused by stressinduced reorientation of interstitial atoms in the octahedral positions of the bcc lattice: these
interactions may include Nb-O, Ti-O, Nb-O-O, Ti-O-O, Nb-N in the -phase.
Since the Nb and Fe are acting as -stabilizers and the -phase is rich in Ti and the -phase is
rich in Nb and, with increasing contents of Nb, the -phase is more stabilized, it may be

suggested that more Nb-O pairs are formed, since as showed in the gas analysis (table 2) these
samples contain a certain amount of oxygen interstitial. Thus, this fact may result in an increase
in the peak intensity. In the subsequent thermal cycles, the precipitation of -phase and/or
growth of -phase occur in each thermal cycle, as observed from XRD patterns. Thus, it is
suggested that firstly the transformation takes place; then, the -phase is formed from
and -phases, consuming -phase particles and some portion of the existing -phase, resulting in
a competition in the growth of and -phases in each thermal cycle and leading to a decrease in
the proportion of -phase and therefore a reduction of the relaxation peak. This statement agrees
with the report by Moffat (Moffat and Larbalestier, 1988), who showed that between 648 K and
698 K the + + transformation can occur in Ti-Nb alloys during aging. Besides, as the
sequence of transformation involves redistribution of elements between regions lean and rich in
solutes, this process may affect the interactions between elements Nb, Ti and interstitial atoms
and the temperature and intensity of the relaxation peak in Ti-20Nb-3Fe and Ti-25Nb-3Fe alloys.

4. Conclusions
The study of anelasticity and the variation of elastic modulus during thermal cycles were
correlated with phase transformations in the Ti-xNb-3Fe (x= 10, 15, 20, 25 wt.%) alloy system.
At low temperatures, signs of a reversible transformation were observed in these alloys.
On the other hand, at higher temperatures, the growth of the metastable -phase was observed
followed by the nucleation and growth of the equilibrium -phase. Thence, in the rapidly
solidified metastable condition, the thermal cycles led from the + microstructure (initial state)
to a + microstructure (final state).
For Ti-20Nb-3Fe and Ti-25Nb-3Fe alloys there were observed two relaxation peak: a peak
around 565 K was associated with stabilization of -phase, since the height was proportional to
the Nb content; the other peak, around 680 K, was linked to -phase, which grows from the phase and decomposition of the metastable -phase. In these alloys, a fall in the relaxation
strength was associated with competition of and -phases, which consume a portion of the
existing -phase. In addition, a behavior of Snoek-type relaxation was identified after these
peaks overlap, to form one peak, which contains components of phase transformations and
interactions of interstitial solutes.
The addition of 3 wt.% Fe, as well as increasing Nb content, induced a fall in the elastic
modulus, since the -phase was more stable. From the E/E curve, retarded growth of -phase

was seen, since the onset temperature for hardening shifts to higher temperatures with increasing
Nb content.
The microstructural changes from the initial rapidly solidified condition were reflected in the
elastic modulus values and hardness. For the alloy with the highest addition of -stabilizing
elements (Ti-25Nb-3Fe), the elastic modulus value was 65 GPa in the rapidly quenched
condition (-phase) and 68 GPa after three thermal cycles (+ phases). In general, Ti-xNb-3Fe
(x= 10, 15, 20, 25 wt.%) alloys, rapidly solidified showed low values of elastic modulus with a
relatively low addition of the isomorphous -stabilizing element (Nb), combined with addition of
the eutectoid -stabilizing element (Fe), relative to the commercial metallic biomedical alloys
(stainless steel, Co-Cr-Mo, CP-Ti and Ti-6Al-4V) employed as biomaterials for orthopedic
implants. This makes Ti-Nb-Fe alloys an attractive system for applications such as metallic
alloys for biomedical implants, in view of the lower melting point and reduced cost associated
with the lower addition of isomorphous -stabilizing element (The noble metal Nb).

Acknowledgments
The authors would like to thank the FAPESP, CNPq and the CAPES for the financial support.

References
Abdel-Hady Gepreel, M., Niinomi, M., 2013. Biocompatibility of Ti-alloys for long-term
implantation. J. Mech. Behav. Biomed. Mater. 20, 407-415.
Afonso, C.R., Ferrandini, P.L., Ramirez, A.J., Caram, R., 2010. High resolution
transmission electron microscopy study of the hardening mechanism through phase separation in
a beta-Ti-35Nb-7Zr-5Ta alloy for implant applications. Acta Biomater. 6, 1625-1629.
Almeida, L.H., Grandini, C.R., Caram, R., 2009. Anelastic spectroscopy in a Ti alloy used
as biomaterial. Mater. Sci. Eng. A 521-22, 59-62.
Baker C, 1971. The Shape-Memory Effect in a Titanium-35 wt.-% Niobium Alloy. Met. Sci.
5, 92-100.
Bertrand, E., Castany, P., Gloriant, T., 2013. Investigation of the martensitic transformation
and the damping behavior of a superelastic TiTaNb alloy. Acta Mater. 61, 511-518.
Boyer R., Collings E.W., Welsch G., 1994. Materials Properties Handbook: Titanium
Alloys. ASM International.
Chaves, J.M., Florncio, O., Silva, P.S., Marques, P.W.B., Schneider, S.G., 2014. Anelastic
relaxation associated to phase transformations and interstitial atoms in the Ti35Nb7Zr alloy. J.
Alloy Comp. 616, 420-425.

Collings, E.W., 1984. The physical metallurgy of titanium alloys. Metals Park, OH :
American Society for Metals, Ohio.
Collings, E.W., 1986. Applied Superconductivity Metallurgy and Physics of Titanium
Alloys. Plenum Press, New York and London.
Elmay, W., Prima, F., Gloriant, T., Bolle, B., Zhong, Y., Patoor, E., Laheurte, P., 2013.
Effects of thermomechanical process on the microstructure and mechanical properties of a fully
martensitic titanium-based biomedical alloy. J. Mech. Behav. Biomed. Mater. 18, 47-56.
Ikehata, H., Nagasako, N., Furuta, T., Fukumoto, A., Miwa, K., Saito, T., 2004. Firstprinciples calculations for development of low elastic modulus Ti alloys. Phys. Rev. B 70,
174113.
Kent, D., Wang, G., Dargusch, M., 2013. Effects of phase stability and processing on the
mechanical properties of Ti-Nb based beta Ti alloys. J. Mech. Behav. Biomed. Mater. 28, 15-25.
Kim, H.Y., Ikehara, Y., Kim, J.I., Hosoda, H., Miyazaki, S., 2006. Martensitic
transformation, shape memory effect and superelasticity of TiNb binary alloys. Acta Mater. 54,
2419-2429.
Lee, C.M., Ho, W.F., Ju, C.P., Chern Lin, J.H., 2002a. Structure and properties of Titanium25 Niobium-x iron alloys. J. Mater. Sci. - Mater. Med. 13, 695-700.
Lee, C.M., Ju, C.P., Chern Lin, J.H., 2002b. Structure-property relationship of cast Ti-Nb
alloys. J. Oral Rehabil. 29, 314-322.
Li, C., Lee, D.-G., Mi, X., Ye, W., Hui, S., Lee, Y., 2013. Phase transformation and age
hardening behavior of new Ti9.2Mo2Fe alloy. J. Alloy Comp. 549, 152-157.
Lin, D.J., Lin, J.H., Ju, C.P., 2002. Structure and properties of Ti-7.5Mo-xFe alloys.
Biomaterials 23, 1723-1730.
Long, M., Rack, H.J., 1998. Titanium alloys in total joint replacement--a materials science
perspective. Biomaterials 19, 1621-1639.
Lopes, E.S.N., Cremasco, A., Afonso, C.R.M., Caram, R., 2011. Effects of double aging
heat treatment on the microstructure, Vickers hardness and elastic modulus of TiNb alloys.
Mater. Charact. 62, 673-680.
Lu, H., Li, C.X., Yin, F.X., Fang, Q.F., Umezawa, O., 2012. Effects of alloying elements on
the Snoek-type relaxation in TiNbXO alloys (X=Al, Sn, Cr, and Mn). Mater. Sci. Eng. A
541, 28-32.
Mantani, Y., Tajima, M., 2006. Phase transformation of quenched martensite by aging in
TiNb alloys. Mater. Sci. Eng. A 438-440, 315-319.
Moffat, D.L., Larbalestier, D.C., 1988. The competition between the alpha and omega phase
in aged Ti-Nb alloys. Metall. Trans. A 19, 1687-1694.

Niinomi, M., 2002. Recent metallic materials for biomedical applications. Metall. Mater.
Trans. A 33, 477-486.
Niinomi, M., 2008. Mechanical biocompatibilities of titanium alloys for biomedical
applications. J. Mech. Behav. Biomed. Mater. 1, 30-42.
Niinomi, M., Nakai, M., 2011. Titanium-Based Biomaterials for Preventing Stress Shielding
between Implant Devices and Bone. Int. J. Biomater. 2011, 836587.
Nowick, A.S., Berry, B.S., 1972. Anelastic Relaxation in Crystalline Solids. Academic
Press, Now York and London.
Ohmori, Y., Ogo, T., Nakai, K., Kobayashi, S., 2001. Effects of -phase precipitation on

, transformations in a metastable titanium alloy. Mater. Sci. Eng. A 312, 182-188.


Reed-Hill, R.E., 1973. Physical Metallurgy Principles. PWS-Kent Publishing, Van Nost.
Reinhold, U. S.
Schaller, R., Fantozzi, G., Gremaud, G., 2001. Mechanical Spectroscopy Q-1 2001. Trans
Tech Publications LTD, Laubisrutistr, CH.
Tane, M., Nakano, T., Kuramoto, S., Niinomi, M., Takesue, N., Nakajima, H., 2013.
Transformation in cold-worked TiNbTaZrO alloys with low body-centered cubic phase
stability and its correlation with their elastic properties. Acta Mater. 61, 139-150.
Tobe, H., Kim, H.Y., Inamura, T., Hosoda, H., Nam, T.H., Miyazaki, S., 2013. Effect of Nb
content on deformation behavior and shape memory properties of TiNb alloys. J. Alloy Comp.
577, S435-S438.
Yin, F., Iwasaki, S., Ping, D., Nagai, K., 2006. Snoek-Type High-Damping Alloys Realized
in -Ti Alloys with High Oxygen Solid Solution. Adv. Mater. 18, 1541-1544.
Zhou, Z.C., Xiong, J.Y., Gu, S.Y., Yang, D.K., Yan, Y.J., Du, J., 2011. Anelastic relaxation
caused by interstitial atoms in -type sintered TiNb alloys. J. Alloy. Comp. 509, 7356-7360.

Figure captions
Fig. 1 - SEM micrographs of rapidly solidified alloys, (a) Ti-10Nb-3Fe (b) Ti-15Nb-3Fe (c) Ti20Nb-3Fe and (d) Ti-25Nb-3Fe, in the as-quenched condition showing a microstructure
becoming more refined with rising Nb content.
Fig. 2 - XRD patterns of (a) Ti-10Nb-3Fe, (b) Ti-15Nb-3Fe, (c) Ti-20Nb-3Fe and (d) Ti-25Nb3Fe rapidly solidified alloys in the as-quenched condition. The inset depict vertically zoomed
portions in a limited 2 range, showing the presence of -phase for samples with lower Nb
content.
Fig 3 - SEM images in BSE (backscattered electrons) mode show the solute distribution in
dendritic microstructure for (a) Ti-10Nb-3Fe and (b) Ti-25Nb-3Fe alloys, with EDS analysis in
the inner region of the dendrites (brighter)) and in the interdendritic region (darker).
Fig. 4 - DSC thermograms of rapidly solidified Ti-xNb-3Fe (x=10, 15, 20 and 25 wt.%) alloys in
the as-quenched condition, showing the phase transformation during two heating cycles at a
heating rate of 1 K/min.
Fig. 5 - Elastic modulus at room temperature plotted against electron concentration (e/a) for the
Ti-xNb-3Fe (), Ti-xNb () and Ti-TM (----) systems.
Fig. 6 - TEM micrographs of as-quenched Ti-10Nb-3Fe sample showing a) bright field (BF) and
b) dark field (DF) images of nanometric -phase precipitate dispersed in -Ti (bcc) grain matrix,
together with respective selected area electron diffraction (SAED) pattern with orientation
relationship of zone axis [2 1 2]//[001] (inset).
Fig. 7 - TEM micrographs of as-quenched Ti-25Nb-3Fe sample of -Ti (bcc) grain boundary
showing a) bright field (BF) and b) dark field (DF) images of nanometric -phase precipitate
dispersed in -Ti matrix, together with respective selected area electron diffraction (SAED)
pattern with orientation relationship of zone axis [1 1 5] (inset).
Fig. 8 - Anelastic relaxation spectra for the alloy Ti-10Nb-3Fe, during (a) first, (b) second and
(c) third heating cycles and (d) XRD patterns showing the variation of phases in each cycle.

Fig. 9 - Anelastic relaxation spectra for the alloy Ti-15Nb-3Fe,


Ti
3Fe, during (a) first, (b) second and
(c) third heating cycles and (d) XRD patterns showing the variation of phases in each cycle.
Fig. 10 - Anelastic relaxation spectra for the alloy Ti
Ti-20Nb-3Fe,
3Fe, during (a) first, (b) second and
(c) third heating cycles and (d) XRD patterns showing the variation of phases in each cycle.
Fig. 11 - Anelastic relaxation spectra for the alloy Ti
Ti-25Nb-3Fe,
3Fe, during (a) first, (b) second and
(c) third heating cycles and (d) XRD patterns showing
showing the variation of phases in each cycle.
Fig. 12 - Dependence of strength of anelastic relaxation peak on Nb content in Ti
Ti-xNb-3Fe (x=
10, 15, 20, 25 wt.%) alloys.

Fig. 1

Fig. 2

Fig. 3

Fig. 4

Fig. 5

[2 1 2]

(-1 -2 1)
(-2 -1 1)
2

(-1 1 0)
1

50 nm

50 nm

Fig. 6

[1 1 5]

50 nm

50 nm

Fig. 7

Fig. 8

Fig. 9

Fig. 10

Fig. 11

Fig. 12

Table captions
Table 1. Chemical analyses for Ti-xNb-3Fe (x= 10, 15, 20, 25 wt.%) alloys.
Table 2. Concentration interstitial solute atoms in the Ti-xNb-3Fe (x= 10, 15, 20, 25 wt.%)
alloys.
Table 3. Elastic modulus (E), Vickers hardness (HV) and phases presents in the alloys Ti-xNb3Fe (x= 10, 15, 20, 25 wt.%) in each thermal cycle.

Table 1.
Nominal
Composition

EDS [wt.%]
Nb

Ti

Ti-10Nb-3Fe

86.9 0.2

Ti-15Nb-3Fe

80.5 0.2

Ti-20Nb-3Fe

75.1 0.2

Ti-25Nb-3Fe

71.3 0.2

Fe

10.2 0.1

2.9 0.1

16.8 0.2

2.7 0.1

21.9 0.2

3.0 0.1

26.0 0.2

2.7 0.1

Table 2.
Alloy

Oxygen [wt.%]

Nitrogen [wt.%]

Ti-10Nb-3Fe

0.15200.0030

0.00320.0001

Ti-15Nb-3Fe

0.19600.0039

0.00380.0001

Ti-20Nb-3Fe

0.16000.0032

0.00270.0001

Ti-25Nb-3Fe

0.18200.0036

0.00370.0001

Table 3.
Ti-10Nb-3Fe

Ti-15Nb-3Fe

Ti-20Nb-3Fe

Ti-25Nb-3Fe

Cycle
number

E
GPa

HV

Phases

E
GPa

HV

Phases

E
GPa

HV

Phases

E
GPa

HV

Phases

As cast

77 6

479 4

71 4

411 5

67 5

326 3

65 6

289 3

95 8

332 3

67 4

314 3

84 6

341 4

69 6

304 3

84 7

331 3

68 4

336 3

82 6

348 4

76 7

310 3

84 7

331 3

67 4

328 3

82 6

354 4

79 7

294 3

Você também pode gostar