You are on page 1of 10

Materials Characterization 172 (2021) 110855

Contents lists available at ScienceDirect

Materials Characterization
journal homepage: www.elsevier.com/locate/matchar

Microstructure globularization of high oxygen concentration dual-phase


extruded Ti alloys via powder metallurgy route
Abdollah Bahador a, *, Junko Umeda a, Hamidreza Ghandvar b, Tuty Asma Abu Bakar b,
Ridvan Yamanoglu c, Ammarueda Issariyapat a, Katsuyoshi Kondoh a
a
JWRI, Osaka University, 11-1 Mihogaoka, Osaka, Ibaraki 567-0047, Japan
b
Faculty of Mechanical Engineering, Universiti Teknologi Malaysia, 81310 UTM Johor Bahru, Johor, Malaysia
c
Kocaeli University, Engineering Faculty, Metallurgical and Materials Engineering Department, Kocaeli, Turkey

A R T I C L E I N F O A B S T R A C T

Keywords: Research related to the substitution of ubiquitous elements for high-cost rare elements during the fabrication of
Microstructure characterization high-performance Ti alloys has attracted significant attention. The microstructure evolution is, however, chal­
α+β titanium lenging when several elements with different properties are used. This research aims to study the effect of Si
Dynamic globularization
addition on the microstructure formation of the dual-phase Ti alloys in the presence of solid solution-forming
Hot extrusion
elements such as Fe and Cu, and a high oxygen concentration. Ti–4Fe–0.5O–3Cu (TFOC) alloys with different
Powder metallurgy
Si contents (0.2, 0.4, and 0.6 wt%) were fabricated from the elemental powders using spark plasma sintering,
followed by hot extrusion. The TFOC sintered alloys with 0.2 and 0.4% Si showed very similar microstructures in
the lamellae α+β dual-phase. The coarsening of α colonies and primary β grains was remarkable when Si content
increased to 0.6%. The obtained primary β grains were, however, significantly smaller than those in previous
studies. In contrast, the refined globular α grains (respective grain sizes = 2.6 and 2.2 μm) embedded in the β
matrix were observed in the hot extruded TFOC–0.2Si and TFOC–0.4Si alloys due to dynamic globularization.
Additionally, EBSD analysis clarified that the hot extruded alloys had the strong fiber textures of (1010) and
(101) for α and β phases respectively, which were consistent with the XRD results. High-resolution TEM ex­
amination demonstrated that Ti2Cu and Ti3Cu intermetallics of a length of 100–400 nm could form preferentially
at α/β interfaces. The experimental results showed that 0.4 wt% Si added to TFOC alloys formed fine globular
microstructures in the presence of hot plastic deformation.

1. Introduction [8]. Zhang et al. [9] showed that Cu improves the antibacterial prop­
erties of Ti alloys, and enhances the strength significantly when Ti2Cu
The α+β titanium alloys form a wide range of high-performance al­ precipitate is formed. Si is added to Ti alloys to increase their creep
loys that are used for different applications, i.e., biomedical and aero­ resistance [10]. However, it is necessary to optimize the content of Si
space industries, due to their high specific strength and excellent and Cu additions, particularly when they are added along with other
corrosion resistance [1,2]. Utilization of β-stabilizing isomorphous re­ alloying elements, to avoid the impairment of ductility due to their
fractory elements (Nb, Ta, and Mo) is well established in the processing higher contents. Using first-principle calculations, Li et al. [11] claimed
of Ti alloys [3–5]. However, due to the comparatively high cost and high that the number of d electrons of transition metals has a significant effect
melting point of this group of alloying elements, ubiquitous elements on attraction or repulsion of Si atoms. Beside, Fe, Cu, and Si are eutec­
(Fe, O, Cu and Si) have gained preference for the fabrication of cost- toid elements and have limited solubility in the α-Ti, thus, triggering the
effective dual phase titanium alloys [6,7]. Fe is known as an effective formation of intermetallic compounds [12]. The solubilities of these
and inexpensive β stabilizer. It was reported that the stabilizing effect of elements in α-Ti are as follows; Fe: maximum 0.5 wt% [13], Cu: less than
1 mass% Fe is equivalent to that of 2.9 mass% Mo [5]. Oxygen with high 1 wt% [14], and Si: Very low solubility of the order of about 1 wt% and
solubility (34 at.%) in α-Ti is a well-known element which is used for 5 wt% in α and β, respectively.
strengthening of Ti alloys via interestitial solid solution strengthening Generally, thermomechanical treatment strongly influences the

* Corresponding author.
E-mail address: abdollah@jwri.osaka-u.ac.jp (A. Bahador).

https://doi.org/10.1016/j.matchar.2020.110855
Received 23 June 2020; Received in revised form 19 October 2020; Accepted 24 December 2020
Available online 26 December 2020
1044-5803/© 2020 Elsevier Inc. All rights reserved.
A. Bahador et al. Materials Characterization 172 (2021) 110855

microstructure of α+β titanium alloys, whereby different grain mor­ 3. Results and discussion
phologies, size, and phase fractions can be obtained. Fine lamellar
morphology is seen when the Ti alloy is deformed in the β phase 3.1. Oxygen–Nitrogen–Hydrogen measurement
following by slow cooling. On the contrary, when a Ti alloy is deformed
in the α+β phase, the microstructure consists of equiaxed morphology The mixed powders successfully consolidated to an almost fully
[15]. The lamellar microstructure is characterized by relatively low dense state and the SPS alloys showed a density of about 4.7 g/cm3
tensile ductility, moderate fatigue properties, and good creep and crack measured by the Archimedes’ principle. Fig. 1 shows the oxygen­
growth resistance. The equiaxed (globular) microstructure has a better –nitrogen–hydrogen content during the fabrication steps. It was found
balance of high strength and ductility at room temperature, as well as that pure Ti as a metal matrix consisted of ~0.29 wt% oxygen, while Fe
fatigue properties, which noticeably depend on the crystallographic and Si contained higher amounts of ~0.34 wt% and ~0.6 wt%,
texture of the hcp α phase [16]. For instance, the yield strength of the respectively. Cu showed the least oxygen content of about 0.05 wt%.
well-known alloy Ti–6Al–4V is improved to 1030 MPa when the Nitrogen and hydrogen contents were lower than 0.03 wt% in all the
microstructure is comprised of fine globular α + β [17]. initial powders, except in Fe which contained 1.1 wt% (Fig. 1a).
Therefore, understanding the effect of hot deformation on the Nevertheless, the volume fractions of these additive elements are
microstructure control and phase fractions is very crucial as it de­ limited, and as illustrated in Fig. 1(b–d), nitrogen and hydrogen
termines the mechanical properties. In this study, as a state-of-art amounts were almost constant after mixing, sintering, and hot extrusion
approach, spark plasma sintering (SPS) was employed to produce (nitrogen: ~0.55 wt%). The oxygen content increased to 0.49 wt% in
Ti–Fe–O–Cu–xSi (x = 0.2, 0.4, and 0.6 wt%). SPS facilitates the fabri­ 0.2Si SPS and 0.4Si SPS, and further to 0.51 wt% in 0.6Si SPS specimen.
cation of highly dense powder metallurgy components with grain The oxygen content did not change after the hot extrusion of 0.2Si SPS
growth control in a short time [18,19]. The sintered alloys were then and 0.4Si SPS, while 0.6Si SPS specimen presented a considerably high
subjected to hot deformation in the α+β phase to obtain the equiaxed oxygen content (0.61 wt%). This high amount of oxygen and Si solutes
α+β microstructure. The resultant microstructures are discussed in the may have affected the hot extrusion in HExt.TFOC-0.6Si alloy.
context of sintering and hot deformation to understand the effect of Si
addition in the presence of other α and β stabilizer elements. 3.2. Microstructure characterization

2. Materials and method The XRD pattern shown in Fig. 2 reveals the microstructure and
existing phases in the sintered and hot extruded specimens. In the full
Three Ti–4Fe–0.5O–3Cu–xSi (x = 0.2, 0.4, and 0.6 wt%) blends were spectra, it is easy to distinguish the high-intensity peaks emitted from
prepared using conventional mechanical milling with ZrO2 balls and a (100)α and (110)β in HExt.TFOC-0.2Si and HExt.TFOC-0.4Si specimens.
ball to powder ratio of 1:5. The average particle sizes of starting powders This is evidence of a strong texture evolution in a direction parallel to
were as follows; Ti: 45 μm, Fe: 5 μm, Cu: 45 μm, Si: 5 μm, and TiO2: 0.5 hot extrusion at lower Si contents. High magnified peaks clearly show
μm (TiO2 was used to add oxygen in the specimens). SPS was employed the decline in the intensity of α and β phases after hot extrusion at (200)
to consolidate the mixed powders using the following parameters: and (211), and (110), (103), and (112), respectively. Additionally, the
heating rate of 20 ◦ C/min, sintering temperature of 1000 ◦ C held for 60 presence of Ti2Cu and Ti3Cu are evidenced by a diffraction peak indi­
min under 30 MPa uniaxial pressure. The sintered specimens were cating (110) and (411) in both sintered and hot extruded specimens
subjected to hot extrusion by pre-heating at 850 ◦ C for 5 min using an containing 0.2Si and 0.4Si, whereas it is not clear when the Si increased
extrusion speed and as extrusion ratio of 6 mm/min and ~18, respec­ to 0.6 wt%.
tively. All the sintered specimens except TFOC–0.6Si were successfully Fig. 3 displays the EBSD imaging of sintered specimens including α
extruded, the reasons for which have been explained in the later sec­ and β phases separately. The sintered samples included colonies of α
tions. The oxygen, nitrogen, and hydrogen contents were measured in all lathes developed inside as well as along the grain boundaries of primary
steps of fabrication for the starting and mixed powders, as well as sin­ β grains. This is a typical microstructure for dual-phase titanium alloys
tered and hot extruded specimens using an Oxygen–Nitrogen–Hydrogen [20]. It is well established that the colony size of α and the primary β
analyzer (EMGA-830, HORIBA). grain size are two main factors in determining mechanical properties of
The produced specimens were examined using X-ray diffraction lamellar α+β phase [21]. Primary β grain boundaries act as nucleation
(XRD, SHIMADZU, XRD-1600) to identify their constituent phases with sites for α phase and directly contribute to both α colony and lath size
Cu-Kα radiation and a scanning speed of 2 deg./min. The electron [22]. Lütjering [23] has shown that the larger the colony size of α, the
backscatter diffraction (EBSD) technique was employed to study the weaker the slip barrier along the primary β grain boundaries, resulting in
crystallographic orientation, phase fraction, and average grain size. For low strength and ductility. Therefore, the EBSD mapping of α and β
EBSD observations, the sintered and hot extruded specimens were phase was investigated in terms of phase fractions, lath thickness, and
conventionally ground with 400 and 1000 grit sandpaper and polished colony of α phase, and the grain size of primary β phase. The α and β
using an 0.05-μm alumina particle solution, followed by vibratory pol­ phase fractions were almost similar in TFOC–0.2Si and TFOC–0.4Si with
ishing (VibroMet® 2, Buehler) for approximately 8 h in a colloidal silica a slightly higher amount of α phase (54 and 55 vol%, respectively).
polishing suspension. The EBSD system included a camera (TSL Digi­ However, the α phase fraction in TFOC–0.6Si remarkably increased up
View IV, EDAX) attached to a field emission scanning microscope to about 73 vol% (Fig. 3). Interestingly, this increment was almost
(FESEM, JSM-6500F, JEOL). A transmission electron microscope (HR- consistent with the α lath thickness and the colony size of α. The α lath
TEM, JEM-2100F, JEOL) operating at 200 kV was employed for high- thickness marginally diminished from 2.7 μm in TFOC–0.2Si to 2.5 μm in
resolution microstructural observations and scanning transmission TFOC–0.4Si, and then increased to 3.8 μm in TFOC–0.6Si (Fig. 4a).
electron microscope (STEM) imaging. The TEM samples were prepared However, the colony size of α increased significantly to 46 μm in the
using a focused ion beam (MultiBeam FIB system, JIB-4500) incorpo­ high Si content specimen, compared to only 25 μm in TFOC–0.2Si and
rated in a scanning electron microscope (SEM) operated at 30 kV. 28 μm in TFOC–0.4Si (Fig. 4b). The average grain size of primary β in the
Microhardness was measured using a Shimadzu (HMV-G) tester by sintered specimens was measured using the intercept method (according
applying a 980.7 mN force for a 10 s dwell time. To verify the accuracy to ASTM E112) and 63 μm, 60 μm, and 65 μm was revealed with respect
of the results, 25 points in a rectangular pattern were measured for each to increasing Si content. Notably, the grain size of the primary β phase
specimen and the average value was considered. was considerably lower than that in the previous studies [24–26] as
shown in Fig. 5. This is associated with the efficient solute drag effect of
multiple alloying elements which lowers primary β grain boundary

2
A. Bahador et al. Materials Characterization 172 (2021) 110855

Fig. 1. Oxygen–nitrogen–hydrogen contents measured during the various steps of the process such as (a) initial powders, (b) after milling (MP: mixed powders), (c)
sintered specimens, and (d) hot extruded rods.

Fig. 2. XRD spectra of sintered (SPS) and hot (HExt) extruded alloys.

mobility and impedes of the grain growth [27].


fg = 1 − exp[ − k.(ε − εc )n ] (1)
Fig. 6 presents the significant influence of hot extrusion at (α+β)
phase region on the microstructure of HExt.TFOC–0.2Si and HExt.
where fg is the volume fraction of dynamic globularization α, εc is the
TFOC–0.4Si composites. It is clearly shown that α lathes have been
critical strain for initiation of dynamic globularization, n is the Avrami-
changed to a globular morphology, and dispersed in the β-phase matrix
exponent, and k is the kinetic constant ad temperature-dependent factor.
due to dynamic globularization. The kinetics of dynamic globularization
However, in the current work, we consider the complete globulari­
has been explained by Song et al. [28] for TC11 dual phase titanium. It
zation of α phase based on the microstructure observations (only limited
has been explained that the deformation temperature and the strain rate
amount of α lathes can be seen in the HExt.TFOC–0.2Si). An average
are critical parameters whereby globularization fraction and globulari­
grain size of 2.6 μm and 2.2 μm was calculated for HExt.TFOC–0.2Si and
zation size can be controlled. The globularization fraction increases with
HExt.TFOC–0.4Si, respectively, using OIM analysis. On the contrary,
strain in a sigmoid way, and it follows the Avrami type equation [29]:

3
A. Bahador et al. Materials Characterization 172 (2021) 110855

attributed to the incomplete hot extrusion.


The separated α and β phase images show almost identical amounts
of α (0.58 vol%) and β (0.42 vol%) phases in all the produced alloys. It is
also evident that grains are oriented dominantly toward <1010> and
〈101〉 in α and β phases, respectively. It implies the formation of a strong
texture and suggests that these directions are parallel to the hot extru­
sion direction. This strong texture formation in lower Si content speci­
mens was in good agreement with the XRD results. In order to further
study the developed textures, pole figures. (PF) and pole plots of HExt.
TFOC–0.4Si was extracted from OIM analysis (Fig. 7). The (1010) PF of α
phase shows a strong intensity in the center of the map along with a
strong ring around. This ring is normally formed due to the crystal
symmetry (Fig. 7a). The PF can be compared to the pole plot, as shown
in Fig. 7b. The large peak at 0 angular distance is due to a strong
alignment of the <1010> crystal directions with the sample normal. The
smaller peak at 60 angular distance is due to ax-symmetrical variants of
<1010> directions. At a first glance, an asymmetric texture at the (101)
PF of β phase is understandable (Fig. 7c). However, this is probably an
indication of an annular texture and it can occur when two sides of the
sample are not parallel. With this consideration, when we smooth and
shift the pole plot, two smaller peaks move to 60 and 90 angular dis­
tances which comply with the standard peaks of (101) fiber texture
(Fig. 7d).
Fig. 8 shows the schematic illustration of microstructure change
supported by FESEM micrograph to understand the mechanism of
globularization. In the initial steps (I, II) acicular α is enlarged by pre-
heating of sintered specimen to α+β region. When the extrusion is
applied α colonies and β grains are slowly oriented with respect to the
applied stress. This process occurs with shearing and kinking of the α
colonies which is governed by the magnitude of the local deformation.
In the next step (III), dislocations of both signs are preferentially
generated along the shear line at the kinking zones. It is followed by
annihilation of opposite sign dislocations due to dislocation cross and
slip which results in the existence of the same sign dislocations [30].
This consequently triggers high energy interface formation at the shear

Fig. 3. IPF mapping, and separated α and β phases demonstrating phase frac­
tions in the SPS alloys; (a) TFOC–0.2Si, (b) TFOC–0.4Si, and (c) TFOC–0.6Si.

HExt.TFOC–0.6Si showed significantly different microstructure as


compared with those with a lower Si content, demonstrating a bimodal
structure comprised of partially globularized and refined α lathes that
are recrystallized in the β grains. This microstructure evolution is Fig. 5. Effect of Fe, B, and Si on the primary β grain size of dual-phase Ti alloys.

Fig. 4. Variation of (a) α lath thickness and (b) α colony size with respect to Si content in SPS specimens.

4
A. Bahador et al. Materials Characterization 172 (2021) 110855

Fig. 6. IPF mapping, and separated α and β phases demonstrating phase fractions in the hot extruded rods; (a) HExt.TFOC–0.2Si, (b) HExt.TFOC–0.4Si, (c) HExt.
TFOC–0.6Si, (d) IPF triangle of HExt.TFOC–0.4Si and (e) schematic illustration of crystallographic orientations in α and β phases.

line (IV). The crystallographic orientation of α and β phases emerged limited amount was also found at the β/β interface (~100 nm length and
with a strong texture that arises from (1010) and (101) planes which are ~20 nm width), as shown in Fig. 9b. The grain boundary precipitation of
known to promote soft deformation in HCP and BCC metals respectively the Ti2Cu intermetallic has already been reported in a previous study by
[31,32]. Finally, continuous applied stress and high energy interfaces Yao et al. [34] on the annealed Ti–Cu alloys. The grain boundaries are
facilitate the migration of interfaces and lowering the energy by forming the main locations where a high density of defects, dislocations, and
the globules (V) [33]. solute concentrations is present. As a result, the grain boundary can be
TEM observations of HExt.TFOC–0.2Si including HR and STEM im­ inferred to be the favored region for the intermetallics to nucleate. On
ages are illustrated in Figs. 9 and 10. The STEM images revealed the the contrary, Cu can easily diffuse into the β matrix, but has low solu­
existence of some precipitations along grain boundaries (Fig. 9a). To bility in α-Ti, due to which, these intermetallic grew preferentially in the
identify their composition, quantitative elemental point scanning using supersaturated α phase.
TEM-attached EDS and HR imaging were performed. Table 1 presents The EDS elemental scanning performed at locations 2 and 3 presents
the EDS obtained from the illustrated points. The point scanning results the quantitative analysis of the alloying elements in the α and β grains.
at locations 1 and 4 were compared to PDF cards (00–055-0296 Cu25.00 Similar to the EDS mapping, these results also show that Fe and Cu
Ti75 at.% and 01–078-9518, Cu33.33 Ti66.67 at.%) and the pre­ dominantly dissolve in the β phase (β stabilizers), while oxygen and Si
cipitations were matched with Ti3Cu and Ti2Cu intermetallics, respec­ diffuse almost identically in both phases.
tively. These precipitations were dominantly located at the α/β grain Fig. 10a shows the HR images of Ti3Cu intermetallics in which L and
interfaces (~100–400 nm length and ~50 nm width). However, their R represent the right (α grain) and left sides (β grain) of the precipitate,
respectively, and M represents the precipitate. Inverse Fast Fourier (FFT)

5
A. Bahador et al. Materials Characterization 172 (2021) 110855

Fig. 7. PF and pole plots of HExt.TFOC–0.4Si, where (a) PF mapping of α, (b) pole plot of α, (c) PF mapping of β, and (d) pole plot of β.

Fig. 8. Schematic illustration of α phase globularization during hot extrusion and FESEM micrograph.

images also are presented for each neighbor grains, where high dislo­ reaction of pre-existing Ti2Cu and β-Ti, according to Canale and Servant
cation densities were revealed in the α phase near the precipitates. The [35]:
lattice fringe measurements in α and β indicate a crystallographic
Ti2 Cu + (β − Ti) ↔ Ti3 Cu (3)
orientation similar to that found in XRD. In contrast, the lattice fringe of
the incoherent precipitate matched with Ti3Cu showed different orien­ Based on the TEM observations, it is assumed that Ti2Cu has partially
tations. The existence of Ti3Cu can be explained as follows. First, Ti2Cu transformed to Ti3Cu due to the short holding time of pre-heating. The
precipitates from β-Ti in the sintering step during cooling by the existence of Ti2Cu was evidenced by the HR image and lattice fringes in
following eutectoid reaction. the inverse FFT image (Fig. 10b).
In addition, Fig. 10c shows the selected area diffraction patterns
β − Ti→α − Ti + Ti2 Cu (2)
(SADPs) corresponding to the locations marked in Fig. 9b. The SADPs of
In the pre-heating of hot extrusion step, Ti3Cu is formed by the α grain are displayed in Fig. 10c (1) as a reference pattern. Fig. 10c (2,3)

6
A. Bahador et al. Materials Characterization 172 (2021) 110855

Fig. 9. STEM imaging and EDS examinations at different locations and interfaces (α/β and β/β) of HExt.TFOC–0.2Si.

shows the SADPs by tilting from α and β grains, respectively. The tilt on content not only stabilizes the ω phase, but also changes its morphology
the α grain was easier to detect from the clear diffraction spots of Ti2Cu from an ellipsoid to an elongated rod during isothermal aging.
intermetallic and capture the related dark field image. The TEM obser­
vations also show that the formation of this intermetallic precipitate is
limited on the β/β interfaces, and is primarily found at the α/β interfaces. 3.3. Microhardness analysis
However, its fraction of the intermetallic is small. It is generally known
that a high density of Ti2Cu can be formed from supersaturated α-Ti via Microhardness evaluation of sintered and hot extruded specimens is
annealing and aging heat treatments of binary Ti–Cu [34,36]. illustrated in Fig. 11. The SPS specimens showed a monotonically
The SADPs obtained from β grain parallel to [113]β zone axis is increasing microhardness with respect to increasing Si solute. However,
shown in Fig. 10c (4). Distinct spots are seen in 1/3 and 2/3 of {112}β, the increment ratio at 0.6 Si was higher than those specimens with a
confirming the presence of ω phase formation in β grains. This generally lower Si content, arising from a difference in microstructure. The α lath
agrees with the previous studies on extruded Ti alloy [37,38]. Addi­ thickness may play the main role in the microhardness of dual phase
tionally, Kilmametov et al. [39] claimed that Ti containing 4 wt% Fe has titanium alloys by hindering the dislocation slip and triggering dislo­
the highest possibility of ω phase development. The presence of this cation pile-ups. As a general rule, the Hall-Petch equation also can be
phase is also evident in the corresponding dark field image. It is neces­ written as following to show this microstructure relation to the micro­
sary to mention that ω phase formation was seen to be limited. Ac­ hardness [42];
cording to literature, oxygen has a suppression effect, particularly when
H = H0 + kH δ− 0.5
(4)
its content is about 0.5 wt% (reported for cold-worked β-Ti alloy) [40].
However, new research reports contradict this hindering effect of oxy­
where, H is microhardness, H0 and kH are constants, and δ is the α lath
gen. For instance, Chou et al. [41] recently claimed that a high oxygen
thickness. However, there is a contradiction in SPS.TFOC-0.6Si which

7
A. Bahador et al. Materials Characterization 172 (2021) 110855

Fig. 10. HR-TEM and inverse FFT images showing (a) Ti3Cu and (b) Ti2Cu intermetallics; (c) diffraction patterns and dark field images obtained from marked areas
in Fig. 9.

might be due to the higher volume fraction of α phase, considering that α


Table 1
phase is harder than β phase.
Elemental EDS point analysis results from Fig. 9a.
Microhardness results did not show a significant difference after hot
Spectrum Ti (at.%) Fe (at. O (at Cu (at Si (at Composition extrusion of low Si content specimens (0.2 and 0.4 wt% Si). On the
%) %) %) %)
contrary, the hardness increased significantly in the alloy with 0.6 wt%
59.80± 3.76 ± 3.45 ± 32.65 ± 0.34 ± Ti3Cu Si which has a bi-modal microstructure and different crystallographic
2.99 0.18 0.17 1.63 0.01
orientations. Although the Hall-Petch effect is considered in the hard­
② 90.46± 0.13 ± 3.55 ± 5.54 ± 0.32 ± Ti-α
4.52 0.0 0.17 0.27 0.01
ness of alloys, the crystallographic orientation of the grains also con­
③ 80.63± 6.73 ± 3.22 ± 9.03 ± 0.39 ± Ti-β tributes effectively. Since a high texture development indicates a low
4.03 0.33 0.16 0.45 0.02 misorientation angle and a nearly-identical slip direction between
④ 68.08± 4.06 ± 4.31 ± 23.0 3± 0.51 ± Ti2Cu adjacent grains, this leads to an easy occurrence of dislocation glide and
3.4 0.2 0.21 1.15 0.02
its movement into the adjacent grain using less force, while a higher

8
A. Bahador et al. Materials Characterization 172 (2021) 110855

Acknowledgments

This work was supported by Project to Create Research and Educa­


tional Hubs for Innovative Manufacturing in Asia, Osaka University, of
the Special Budget Project of the Ministry of Education, Culture, Sports,
Science, and Technology. A part of this study was also financially sup­
ported by Cross-ministerial Strategic Innovation Promotion Program
(SIP), “Materials Integration for revolutionary design system of struc­
tural materials” (Funding agency: JST). The authors would like to
specially thank Mr. Takeshi Murakami for his technical support during
TEM observations.

References

[1] I. Yadroitsev, P. Krakhmalev, I. Yadroitsava, Selective laser melting of Ti6Al4V


alloy for biomedical applications: temperature monitoring and microstructural
evolution, J. Alloys Compd. 583 (2014) 404–409, https://doi.org/10.1016/j.
Fig. 11. Variation of microhardness of SPS and HExt. TFOC alloys with respect
jallcom.2013.08.183.
to Si content. [2] K. Kapoor, P. Ravi, D. Naragani, J. Park, J.D. Almer, M.D. Sangid, Materials
Characterization Strain rate sensitivity, microstructure variations, and stress-
assisted β → α phase transformation investigation on the mechanical behavior of
force is needed for a dislocation to cross the grain boundary in differ­
dual-phase titanium alloys, Mater. Charact. 166 (2020), 110410, https://doi.org/
ently textured adjacent grains [43]. Therefore, HExt.TFOC-0.2Si and 10.1016/j.matchar.2020.110410.
HExt.TFOC-0.4Si with highly textured equiaxed microstructures have [3] V. Brailovski, S. Prokoshkin, M. Gauthier, K. Inaekyan, S. Dubinskiy, M. Petrzhik,
lower microhardness than the HExt.TFOC-0.6Si with randomly textured M. Filonov, Bulk and porous metastable beta Ti-Nb-Zr(Ta) alloys for biomedical
applications, Mater. Sci. Eng. C 31 (2011) 643–657, https://doi.org/10.1016/j.
and bi-modal microstructure. In addition, high oxygen content of HExt. msec.2010.12.008.
TFOC-0.6Si may also increase the microhardness. [4] A. Bahador, S. Kariya, J. Umeda, E. Hamzah, K. Kondoh, Tailoring microstructure
and properties of a superelastic Ti–Ta alloy by incorporating spark plasma sintering
with thermomechanical processing, J. Mater. Eng. Perform. 28 (2019) 3012–3020,
4. Conclusions https://doi.org/10.1007/s11665-019-04061-8.
[5] X.H. Min, K. Tsuzaki, S. Emura, K. Tsuchiya, Enhancement of uniform elongation in
Three different blends of pre-mixed Ti–4Fe–5O–3Cu–xSi (x = 0.2, high strength Ti-Mo based alloys by combination of deformation modes, Mater. Sci.
Eng. A 528 (2011) 4569–4578, https://doi.org/10.1016/j.msea.2011.02.071.
0.4, and 0.6 wt%) powder were densified using SPS process and pro­ [6] D.V. Louzguine, H. Kato, A. Inoue, High strength and ductile binary Ti-Fe
cessed further by hot extrusion. The aim was to fabricate cost-effective composite alloy, J. Alloys Compd. 384 (2004) 10–12, https://doi.org/10.1016/j.
dual-phase titanium alloys and study the effect of Si addition on jallcom.2004.03.114.
[7] M. Kikuchi, Y. Takada, S. Kiyosue, M. Yoda, M. Woldu, Z. Cai, O. Okuno, T. Okabe,
microstructure, phase fraction, and morphology of grains. The sintered
Mechanical properties and microstructures of cast Ti-Cu alloys, Dent. Mater. 19
specimens showed dual-phase (α+β) titanium, where α lathes were (2003) 375–381, https://doi.org/10.1016/S0109-5641(02)00080-5.
located inside and along the primary β grain boundaries. The Si addition [8] B. Sun, S. Li, H. Imai, T. Mimoto, J. Umeda, K. Kondoh, Fabrication of high-
strength Ti materials by in-process solid solution strengthening of oxygen via P/M
resulted in slightly coarsening of primary β grains, especially when it
methods, Mater. Sci. Eng. A 563 (2013) 95–100, https://doi.org/10.1016/j.
increased to 0.6 wt%. The hot extrusion at the temperature of dual-phase msea.2012.11.058.
stability successfully altered the lamellar grain morphology of sintered [9] E. Zhang, X. Wang, M. Chen, B. Hou, Effect of the existing form of cu element on
alloys to refined globular shaped with a grain size of 2.6 and 2.2 μm in the mechanical properties, bio-corrosion and antibacterial properties of Ti-Cu
alloys for biomedical application, Mater. Sci. Eng. C 69 (2016) 1210–1221, https://
HExt.TFOC–0.2Si and HExt.TFOC–0.4Si, respectively. The α and β phase doi.org/10.1016/j.msec.2016.08.033.
fractions were almost similar after hot extrusion in all the alloys with the [10] W. Jia, W. Zeng, Y. Zhou, J. Liu, Q. Wang, High-temperature deformation behavior
ratio of α/β: ~6/4. Additionally, EBSD analysis clarified the strong of Ti60 titanium alloy, Mater. Sci. Eng. A 528 (2011) 4068–4074, https://doi.org/
10.1016/j.msea.2011.01.113.
crystallography alignment in both α and β phases of hot extruded [11] Y. Li, Y. Chen, J.R. Liu, Q.M. Hu, R. Yang, Cooperative effect of silicon and other
specimen resulted in the development of (1010) α and (101) β fiber alloying elements on creep resistance of titanium alloys: insight from first-
textures, which agreed with the XRD characterization. TEM in­ principles calculations, Sci. Rep. 6 (2016) 1–8, https://doi.org/10.1038/
srep30611.
vestigations revealed the existence of Ti2Cu and Ti3Cu intermetallics [12] R.P. Kolli, W.J. Joost, S. Ankem, Phase stability and stress-induced transformations
which preferentially formed along α/β interfaces. Si and O were found to in beta titanium alloys, JOM 67 (2015) 1273–1280, https://doi.org/10.1007/
diffuse homogenously in both α and β phase while Fe and Cu mainly s11837-015-1411-y.
[13] A.S. Gornakova, B.B. Straumal, A.N. Nekrasov, A. Kilmametov, N.S. Afonikova,
diffused into the β phase. These results showed a successful control of Grain boundary wetting by a second solid phase in Ti-Fe alloys, J. Mater. Eng.
the microstructure in achieving fine equiaxed grains of α+β dual-phase Perform. 27 (2018) 4989–4992, https://doi.org/10.1007/s11665-018-3300-3.
when the Si content did not exceed 0.4 wt% in the presence of a high [14] S.Y. Chang, L.C. Tsao, Y.H. Lei, S.M. Mao, C.H. Huang, Brazing of 6061 aluminum
alloy/Ti-6Al-4V using Al-Si-Cu-Ge filler metals, J. Mater. Process. Technol. 212
concentration of oxygen. The ability to control the microstructure has (2012) 8–14, https://doi.org/10.1016/j.jmatprotec.2011.07.014.
implications on the strengthening of dual-phase titanium alloys using [15] F.W. Syed, V. Anil Kumar, R.K. Gupta, A.K. Kanjarla, Role of microstructure on the
ubiquitous elements. In future work, mechanical properties, including tension/compression asymmetry in a two-phase Ti-5Al-3Mo-1.5V titanium alloy,
J. Alloys Compd. 795 (2019) 151–162, https://doi.org/10.1016/j.
strengthening and deformation mechanisms based on grain size and
jallcom.2019.04.272.
solid solution effects, will be discussed. Additionally, considering the [16] J. Sieniawski, W. Ziaja, K. Kubiak, M. Motyk, Microstructure and mechanical
potential for high-temperature applications, the stabiliy of the alloys at properties of high strength two-phase titanium alloys, Titan. Alloy. - Adv. Prop.
Control. (2013), https://doi.org/10.5772/56197.
high temperatures will be studied using in-situ SEM observations.
[17] M. Motyka, K. Kubiak, J. Sieniawski, W. Ziaja, Phase transformations and
characterization of α + β titanium alloys, Compr. Mater. Process. 2 (2014) 7–36,
Declaration of Competing Interest https://doi.org/10.1016/B978-0-08-096532-1.00202-8.
[18] A. Bahador, E. Hamzah, K. Kondoh, T.A. Abu Bakar, F. Yusof, H. Imai, S.N. Saud,
M.K. Ibrahim, Effect of deformation on the microstructure, transformation
The authors declare that they have no known competing financial temperature and superelasticity of Ti-23 at% Nb shape-memory alloys, Mater. Des.
interests or personal relationships that could have appeared to influence 118 (2017) 152–162, https://doi.org/10.1016/j.matdes.2016.12.048.
the work reported in this paper. [19] A. Bahador, E. Hamzah, K. Kondoh, T. Asma Abubakar, F. Yusof, J. Umeda, S.
N. Saud, M.K. Ibrahim, Microstructure and superelastic properties of free forged
Ti–Ni shape-memory alloy, Trans. Nonferrous Metals Soc. China 28 (2018)
502–514, https://doi.org/10.1016/S1003-6326(18)64683-7.

9
A. Bahador et al. Materials Characterization 172 (2021) 110855

[20] R. Yamanoglu, E. Efendi, F. Kolayli, H. Uzuner, I. Daoud, Production and [32] M. Zhang, J. Zhang, D.L. McDowell, Microstructure-based crystal plasticity
mechanical properties of Ti-5Al-2.5Fe-xCu alloys for biomedical applications, modeling of cyclic deformation of Ti-6Al-4V, Int. J. Plast. 23 (2007) 1328–1348,
Biomed. Mater. 13 (2018), https://doi.org/10.1088/1748-605X/aa957d. https://doi.org/10.1016/j.ijplas.2006.11.009.
[21] D. Jeong, Y. Kwon, M. Goto, S. Kim, High cycle fatigue and fatigue crack [33] C.H. Park, K.T. Park, D.H. Shin, C.S. Lee, Microstructural mechanisms during
propagation behaviors of β-annealed Ti-6Al-4V alloy, Int. J. Mech. Mater. Eng. 12 dynamic globularization of Ti-6A1-4V alloy, Mater. Trans. 49 (2008) 2196–2200,
(2017) 1–10, https://doi.org/10.1186/s40712-016-0069-8. https://doi.org/10.2320/matertrans.L-MRA2008832.
[22] J. Tiley, T. Searles, E. Lee, S. Kar, R. Banerjee, J.C. Russ, H.L. Fraser, Quantification [34] X. Yao, Q.Y. Sun, L. Xiao, J. Sun, Effect of Ti2Cu precipitates on mechanical
of microstructural features in α/β titanium alloys, Mater. Sci. Eng. A 372 (2004) behavior of Ti-2.5Cu alloy subjected to different heat treatments, J. Alloys Compd.
191–198, https://doi.org/10.1016/j.msea.2003.12.008. 484 (2009) 196–202, https://doi.org/10.1016/j.jallcom.2009.04.095.
[23] G. Lütjering, Influence of processing on microstructure and mechanical properties [35] P. Canale, C. Servant, Thermodynamic assessment of the Cu-Ti system taking into
of (α + β) titanium alloys, Mater. Sci. Eng. A 243 (1998) 32–45, https://doi.org/ account the new stable phase CuTi3, Zeitschrift Fuer Met. Res. Adv. Tech. 93
10.1016/s0921-5093(97)00778-8. (2002) 273–276, https://doi.org/10.3139/146.020273.
[24] M.J. Bermingham, S.D. McDonald, M.S. Dargusch, D.H. StJohn, The mechanism of [36] Z. Lincai, D. Xiaoming, Y. Wei, Z. Man, S. Zhenya, Effect of prestrain on
grain refinement of titanium by silicon, Scr. Mater. 58 (2008) 1050–1053, https:// precipitation behaviors of Ti-2.5Cu alloy, High Temp. Mater. Process. 37 (2018)
doi.org/10.1016/j.scriptamat.2008.01.041. 487–493, https://doi.org/10.1515/htmp-2017-0006.
[25] S. Tamirisakandala, R.B. Bhat, J.S. Tiley, D.B. Miracle, Grain refinement of cast [37] R.J. Talling, R.J. Dashwood, M. Jackson, D. Dye, On the mechanism of
titanium alloys via trace boron addition, Scr. Mater. 53 (2005) 1421–1426, https:// superelasticity in gum metal, Acta Mater. 57 (2009) 1188–1198, https://doi.org/
doi.org/10.1016/j.scriptamat.2005.08.020. 10.1016/j.actamat.2008.11.013.
[26] Y. Liu, L.F. Chen, H.P. Tang, C.T. Liu, B. Liu, B.Y. Huang, Design of powder [38] J. Coakley, V.A. Vorontsov, K.C. Littrell, R.K. Heenan, M. Ohnuma, N.G. Jones,
metallurgy titanium alloys and composites, Mater. Sci. Eng. A 418 (2006) 25–35, D. Dye, Nanoprecipitation in a beta-titanium alloy, J. Alloys Compd. 623 (2015)
https://doi.org/10.1016/j.msea.2005.10.057. 146–156, https://doi.org/10.1016/j.jallcom.2014.10.038.
[27] J. Shen, B. Chen, J. Umeda, J. Zhang, Y. Li, K. Kondoh, Rate sensitivity and work- [39] A.R. Kilmametov, Y. Ivanisenko, A.A. Mazilkin, B.B. Straumal, A.S. Gornakova, O.
hardening behavior of an advanced Ti-Al-N alloy under uniaxial tensile loading, B. Fabrichnaya, M.J. Kriegel, D. Rafaja, H. Hahn, The α→ω and β→ω phase
Mater. Sci. Eng. A 744 (2019) 630–637, https://doi.org/10.1016/j. transformations in Ti–Fe alloys under high-pressure torsion, Acta Mater. 144
msea.2018.12.066. (2018) 337–351, https://doi.org/10.1016/j.actamat.2017.10.051.
[28] H.W. Song, S.H. Zhang, M. Cheng, Dynamic globularization kinetics during hot [40] M. Tane, T. Nakano, S. Kuramoto, M. Niinomi, N. Takesue, H. Nakajima, ω
working of a two phase titanium alloy with a colony alpha microstructure, J. Alloys Transformation in cold-worked Ti-Nb-Ta-Zr-O alloys with low body-centered cubic
Compd. 480 (2009) 922–927, https://doi.org/10.1016/j.jallcom.2009.02.059. phase stability and its correlation with their elastic properties, Acta Mater. 61
[29] X. Ma, W. Zeng, F. Tian, Y. Zhou, The kinetics of dynamic globularization during (2013) 139–150, https://doi.org/10.1016/j.actamat.2012.09.041.
hot working of a two phase titanium alloy with starting lamellar microstructure, [41] K. Chou, E.A. Marquis, Oxygen effects on ω and α phase transformations in a
Mater. Sci. Eng. A 548 (2012) 6–11, https://doi.org/10.1016/j.msea.2012.03.022. metastable β Ti–Nb alloy, Acta Mater. 181 (2019) 367–376, https://doi.org/
[30] T. Seshacharyulu, S.C. Medeiros, J.T. Morgan, J.C. Malas, W.G. Frazier, Y.V.R. 10.1016/j.actamat.2019.09.049.
K. Prasad, Hot deformation and microstructural damage mechanisms in extra-low [42] H. Galarraga, R.J. Warren, D.A. Lados, R.R. Dehoff, M.M. Kirka, P. Nandwana,
interstitial (ELI) grade Ti-6Al-4V, Mater. Sci. Eng. A 279 (2000) 289–299, https:// Effects of heat treatments on microstructure and properties of Ti-6Al-4V ELI alloy
doi.org/10.1016/S0921-5093(99)00173-2. fabricated by electron beam melting (EBM), Mater. Sci. Eng. A 685 (2017)
[31] W.B. Hutchinson, M.R. Barnett, Effective values of critical resolved shear stress for 417–428, https://doi.org/10.1016/j.msea.2017.01.019.
slip in polycrystalline magnesium and other hcp metals, Scr. Mater. 63 (2010) [43] A. Bahador, J. Umeda, S. Tsutsumi, E. Hamzah, F. Yusof, H. Fujii, K. Kondoh,
737–740, https://doi.org/10.1016/j.scriptamat.2010.05.047. Asymmetric local strain, microstructure and superelasticity of friction stir welded
Nitinol alloy, Mater. Sci. Eng. A 767 (2019) 138344, https://doi.org/10.1016/j.
msea.2019.138344.

10

You might also like