You are on page 1of 25

Journal Pre-proof

An investigation of the tensile deformation mechanism of a high-oxygen Ti2448 alloy


fabricated by powder metallurgy by use of the in-situ EBSD method

Xia Li, Jinhui Wang, Shulong Ye, Yinghao Zhou, Peng Yu

PII: S0921-5093(21)01247-8
DOI: https://doi.org/10.1016/j.msea.2021.141981
Reference: MSA 141981

To appear in: Materials Science & Engineering A

Received Date: 3 February 2021


Revised Date: 30 July 2021
Accepted Date: 24 August 2021

Please cite this article as: X. Li, J. Wang, S. Ye, Y. Zhou, P. Yu, An investigation of the tensile
deformation mechanism of a high-oxygen Ti2448 alloy fabricated by powder metallurgy by use of
the in-situ EBSD method, Materials Science & Engineering A (2021), doi: https://doi.org/10.1016/
j.msea.2021.141981.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. 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.

© 2021 Published by Elsevier B.V.


CRediT authorship contribution statement

Xia Li: Conceptualization; Investigation; Roles/Writing - original draft; Writing -

review & editing. Jinhui Wang: Conceptualization; Data curation. Shulong Ye:

Formal analysis; Funding acquisition. Yinghao Zhou: Data curation; Methodology.

Peng Yu: Project administration; Supervision.

f
r oo
-p
re
lP
na
ur
Jo
An investigation of the tensile deformation mechanism of a high-oxygen

Ti2448 alloy fabricated by powder metallurgy by use of the in-situ EBSD

method

Xia Li1, Jinhui Wang2, Shulong Ye1, Yinghao Zhou1 and Peng Yu1,*
1
Department of Materials Science and Engineering, Southern University of Science and

Technology, Shenzhen, 518055, P. R. China


2
Qinghai Provincial Key Laboratory of New Light Alloys, Qinghai Provincial Engineering

of
Research Center of High-Performance Light Metal Alloys and Forming, Qinghai University,

ro
Xining, 810016, PR China

-p
re
Abstract: Ti2448 alloy fabricated by powder metallurgy using elemental powders of Ti, Nb, Zr
lP

and Sn exhibits good ductility but a high content of oxygen, and very limited work hardening in
na

the tensile test. Its deformation mechanism is investigated by in-situ EBSD test conducted in an
ur

SEM equipped with a dynamic tensile stage. The results indicate that both dislocation slide and
Jo

stress-induced phase transformation contribute to the plastic deformation of the alloy. Subsequent

to yielding (after 3% of strain), β phase grains with preferential orientation 〈113〉𝛽 tend to be

transformed into α" martensite phase, which significantly reduces the plastic deformation caused

by the dislocation slide and therefore leads to low work hardening.

Keywords: Deformation mechanism; Work hardening; Martensite transformation; Powder

metallurgy; Titanium alloy

*Corresponding author, E-mail address: yup@sustech.edu.cn, yp1975@gmail.com (Peng Yu)

1
1. Introduction

Beta-titanium alloys, which have elastic modulus compatible with that of the human bones, are

regarded as excellent materials for medical applications. Since the materials are used as implants

which need to serve in human bodies for life time, reliable mechanical properties became a crucial

issue. Many studies[1, 2] have shown that the mechanical properties of β-Ti alloys are greatly

influenced by their microstructures. Therefore, a better understanding of the microstructure

transformation during deformation of the β-Ti alloys is essential for predicting the mechanical

of
properties of the alloys. In order to do so, various advanced technologies, such as in-situ

ro
synchrotron X-ray diffraction[3, 4], in-situ scanning and transmission electron microscopy[5], and

-p
digital image correlation (DIC) method for the assessment of displacement and strain fields[6] are
re
adopted in previous researches.
lP
na

Recently, a β-Ti alloy invented by Hao et. al. has attracted many researchers’ attentions. With a

nominal composition of Ti-24Nb-4Zr-8Sn(wt.%)[7], it is widely known as Ti2448 alloy. This


ur

material has an extremely low elastic modulus, therefore being regarded as a representative of the
Jo

next generation of medical materials. Its mechanical properties have been investigated intensively.

Hao et al.[7] found the super-elastic behavior of the Ti2248 alloy fabricated by hot-forging and

hot-rolling; and attributed the super-elasticity to the partially reversible process of stress-induced

phase transformation. Yang et al.[8] investigated the deformation mechanisms of the super-

elasticity in the Ti2448 alloy with very low oxygen content, which was fabricated by hot-forging

and cold-rolling. The results suggested that the super-elasticity derived from the hierarchal

martensitic twinning and large deformations resulting from the forging and rolling processes. Fer

et al.[9] used Spark Plasma Sintering (SPS) to fabricate Ti2448 alloy with a heterogeneous

2
microstructure, which consists of β coarse grain regions and α/β dual phase ultra-fine grain regions.

The relationship between the mechanical behavior and the heterogeneous microstructure was also

discussed.

In the meanwhile, increasing demand for implants and surgical instruments with extremely

complex morphology required the use of net shape fabrication techniques in the fabrication of

medical Ti alloys. There have been researches focused on processing Ti2448 alloy by selective

of
laser melting (SLM)[10-12], metal injection molding (MIM)[13] and electron beam melting

ro
(EBM)[14-16]. Powders are used as raw materials in these methods, which have large surface

-p
areas and high oxygen concentrations. As a result, the oxygen content of products fabricated by
re
these net-shape fabrication processes (such as powder metallurgy (PM), MIM, SLM and EBM) is
lP

relatively higher than those produced by traditional processes (such as forging and casting).
na

Many researches indicate that oxygen influences the mechanical properties of β-Ti alloys. It can
ur

destabilize the β phase in Ti-Mo alloy, Ti-Nb alloy and Ti-23Nb-0.7Ta-2Zr alloys [17-19]. Besse
Jo

et al.[20] found that oxygen tends to hinder the martensite transformation in Ti–23Nb–0.7Ta–2Zr

(TNTZ) titanium alloy. Obbard et al.[21] studied the effect of oxygen on martensite transformation

and super-elasticity of forged Ti2448 alloy. They found that oxygen reduces the non-line

recoverable strain, shortens the stress plateau and obscures the double yield points. In addition,

Dai et al.[22] studied the effect of oxygen on Ti2448 alloy via first principles calculations. Results

showed that oxygen has a weak effect on elastic modulus, but is effective in changing phase

stability. Therefore, it is expected that the high-oxygen β-Ti alloys produced by net-shape

fabrication methods and the low-oxygen β-Ti alloys fabricated by conventional methods would

3
have differences in mechanical properties and deformation mechanisms. However, there is no

systematic research on the deformation mechanism of those high-oxygen β-Ti alloys.

In our previous researches, Ti2448 alloy was fabricated by powder metallurgy (PM) [23] .

Although it has a high content of oxygen (3400 to 3800 ppm), we have proved that the alloy with

a single β phase can be obtained through solution treatment at proper temperatures and subsequent

quenching. The alloy exhibits mechanical properties (ultimate tensile strength: 725±14 MPa; yield

of
strength: 655±12 MPa; elongation: 19±1%) even comparable to those fabricated by forging and

ro
rolling. In this study, this alloy is used as a prototype in an in-situ tensile test to reveal the

-p
deformation mechanism of high-oxygen Ti2448 alloy (~3470 ppm).
re
lP

2. Material and methods


na

Ti2448 alloy with a nominal composition of Ti-24Nb-4Zr-8Sn (wt.%) was manufactured by using
ur

elemental powders of Ti (spherical shape, D50 (medium diameter) = 35 μm), Nb (irregular shape,
Jo

D50 = 16 μm), Zr (irregular shape, D50 = 6 μm) and Sn (near spherical shape, D50 = 12 μm).

These powders were mixed in a three-dimension mixer (Turbula T2F, Switzerland) and cold-

pressed in a pressing machine (MTI Corporation, YLJ-60T, China) to fabricate green compacts of

tensile bars. Subsequently, Ti2448 alloy samples were fabricated according to a procedure

optimized in our previous researches [23, 24]. In this process, the green compacts were sintered at

1400 ℃ under flowing Ar in a tube furnace (HF-kejing, GSL-1600X, China). The as-sintered

samples were solution-treated at 980 ℃ for 1 hour and subsequently water-quenched. The tensile

properties of the as-fabricated Ti2448 alloy were tested by a universal testing machine (MTI

Corporation, YLJ-60T, China).

4
In order to investigate the deformation mechanism of the Ti2448 alloy, in-situ tensile tests were

carried out on the alloy in a field emission scanning electron microscope (SEM, Zeiss 6035)

equipped with an electron back-scattered diffraction (EBSD, Oxford Instruments, NordlysMax2)

detector and a dynamic tensile stage. When the sample was deformed on the dynamic tensile stage,

information about the phase compositions and grain orientations of the sample were in-situ

collected by the EBSD detector. Later, the results were analyzed using a professional software

of
(Oxford HKL Channel 5), which gave clues on the deformation mechanism of the alloy.

ro
-p
The phase compositions of the alloy were determined by an X-ray diffractometer (XRD, Bruker
re
AXS D8). The surface morphologies of samples before and after the tension test were observed by
lP

an Atomic Force Microscope (AFM, Asylum Research, MFP-3D-Stand Alone). TEM samples
na

were prepared by electrolytic polishing using a double spray (Struers, 05396133) and investigated

by a transmission electron microscopy (TEM, Tecnai F30).


ur
Jo

3. Results

Fig. 1 (a) shows the XRD pattern of the quenched sample before and after the tensile test, in which

only peaks of β phase (bcc) are revealed. The result indicates that Ti2448 alloy with a single β

phase has been fabricated by powder metallurgy. In fact, this phenomenon has been discussed in

our previous study [23, 24]. Through solution treatment and subsequent water quenching, the

precipitation of α phase in Ti2448 alloy can be successfully suppressed, producing samples with a

monolithic β phase. Hence, due to the absence of acicular a phase, the water-quenched samples

exhibit good mechanical properties. Fig. 1 (b) shows the true and engineering stress-strain curves

5
of the tensile test of the quenched sample obtained, which reveal that the material has good

mechanical properties. The ultimate tensile strength (UTS), yielding stress (σ0), strain-at-fracture

(εf) and Young’s modulus (E) are 725±14 MPa, 655±12 MPa, 19±1% and 57.2±1 GPa,

respectively. However, the engineering stress-strain curve shows a long stress plateau where stress

varies little as the plastic deformation goes on, implying that work hardening, a phenomenon

widely found in plastic deformation of metals, is lacking in the Ti2448 alloy of this research. In

order to rule out the influences of the variation of the cross-section area in plastic deformation on

of
the stress calculation, a true stress-strain curve is also plotted. And shown in Fig. 1 (b). The true

ro
stress increases slightly in the plastic deformation stage. The strain hardening ratio, which is

-p
defined as the ratio of the ultimate tensile strength (UTS) to the yielding strength (YS) of a material,
re
is only 1.10 for Ti2448 alloy in this research. Table 1 lists the UTS, YS and strain hardening ratios
lP

of β-Ti alloys reported in previous researches[25-28]. The results indicate that the strain hardening
na

ratio of the Ti2448 alloy of our research is lower than most of the β-Ti alloys, except Ti-35Nb-

7Zr-5Ta-0.46O and Ti-35Nb-7Zr-5Ta-0.68O[27]. The latter two alloys are similar to the Ti2448
ur

alloy in this research in which they also have high levels of oxygen. Both Ti2448 alloy (~3470
Jo

ppm) in this work and Ti-Nb-Zr-Ta alloy have high oxygen content, which leads to insignificant

work hardening.

6
of
Fig. 1 (a) XRD spectrum of the Ti2448 alloy sample fabricated by powder metallurgy, (b) true and

ro
engineering stress-strain curves of the sample obtained from tensile tests, (c) SEM images of the

-p
fracture surfaces and (d) (e) (f) (g) AFM images of the sample: (d) (e) before the tensile test and
re
(f) (g) after the tensile test.
lP
na

Table 1 Summary of strain hardening ratio of β-titanium alloys of previous researches [24-27].
(Strain hardening ratio = UTS/YS, where UTS and YS are the ultimate tensile strength and the
ur

yielding strength.)
Alloys YS/MPa UTS/MPa UTS/YS
Jo

This work 655±13 725±14 1.10


Ti-30Nb[28] 500 700 1.4
Ti-15Mo[25] 544 874 1.60
Ti-30Ta[28] 590 740 1.25
Ti-29Nb-13Ta-4.6Zr[26] 400 1025±25 2.56
Ti-35Nb-7Zr-5Ta-0.46O[27] 806 929 1.15
Ti-35Nb-7Zr-5Ta-0.68O[27] 1036 1180 1.13

The surface of the quenched sample is observed by SEM after the tensile test. The image is shown

in Fig. 1(c), which shows slip bands along different directions. This observation is further

confirmed by AFM. Fig. 1 (d) and (e) shown 2-D and 3-D AFM images of the Ti2448 alloy prior

7
to the tensile test, which reveal a smooth surface without any slip bands. In the meanwhile, Fig. 1

(f) and (g) show the AFM images of the Ti2448 alloy subsequent to the tensile test. Grooves with

a depth ranging from 800 to 900 nm can be clearly seen. These results indicate that dislocation slip

does occur in the plastic deformation of Ti2448 alloy and accounts for the slight work hardening

of the alloy.

TEM is used to investigate the deformation mechanism of the Ti2448 alloy. Fig. 2(a) shows a

of
TEM bright field (BF) image of the Ti2448 alloy before tensile test. Fig. 2(b) shows a selected

ro
area electron diffraction (SAED) pattern of the same sample, which is identified as the β phase. In

-p
comparison, Fig. 2(c), (d) and (e) show the BF image, dark field (DF) image and SAED pattern of
re
the sample after the tensile test. Fig. 2(c) shows the structure of the dislocations in the deformed
lP

Ti2448 alloy. The dislocations intertwine and form an entangled structure, which can serve as a
na

barrier to ongoing motion of dislocations and leads to work hardening of the alloy in deformation.

Apart from the β phase, a new phase, although in the trace amount, has been revealed by TEM in
ur

the deformed Ti2448 alloy. Dark field image from the α" reflection highlighted are presented in
Jo

Fig. 2(d). The new phase appears as a bright stripe between the two red broken lines as labeled in

Fig. 2(d). The SAED pattern (Fig. 2(e)) is obtained from the sample after the tensile test, which is

used to directly determine the orientation relationship of α" and β phases. The electron beam is

incident symmetrically with respect to the two phases, i.e., the zone axes of the two phases are

both parallel to the incident beam. In Fig. 2(e), two sets of diffraction spots are separated from the

SAED pattern, which are identified as orthorhombic α" and bcc β phases with specific crystal

indices. Hence, parallel zone axes of [001]𝛽 and [1̅10]𝛼 " are determined. In addition, two

reciprocal vectors are found parallel to each other in the SAED pattern, in which the crystal indices

8
are (11̅0)𝛽 and (110)𝛼" , respectively. Therefore, it is analyzed that the α" phase has a specific

orientation relationship with respect to the β phase, i.e., [001]𝛽 //[1̅10]𝛼 " , (11̅0)𝛽 //(110)𝛼 " .

This unexpected observation suggests that martensite transformation from β to α" phase occurs in

the plastic deformation of the Ti2448 alloy. Therefore, its deformation mechanism cannot be

attributed to the dislocation slip along.

of
ro
-p
re
lP
na
ur
Jo

Fig. 2 (a) TEM image of the Ti2448 alloy sample before the tensile test and (b) corresponding

SAED pattern indicating that the sample has a single β phase; (c) bright-field TEM image, (d)

dark-field TEM image and (e) corresponding SAED pattern of the sample after the tensile test

indicating that α" phase exists in the matrix of the β phase having an specific orientation

relationship with respect to the matrix, i.e., [001]𝛽 //[1̅10]𝛼 " , (11̅0)𝛽 //(110)𝛼" . (Fig. 2(e) is

SAED pattern obtained by reverse phase processing of PS in order to show the α" phase more

clearly.)

4. Discussion

9
In order to confirm the above claim, in situ EBSD experiment is further carried out on the Ti2448

alloy in an SEM equipped with a dynamic tensile stage. The experiment is illustrated in Fig. 3(a).

When the sample is stretched in the tensile stage, the secondary electrons and back scattering

electrons generated by electron beam of the SEM are in-situ collected by the detectors. The results

are used to analyze the change of microstructures of the sample during deformation. The stress-

strain curve obtained from the in-situ test is shown in Fig. 3 (b). The curve highly resembles the

one shown in Fig. 1(b), except that a few stress troughs can be found in the curve obtained from

of
the in-situ test. Those troughs are caused by stress relaxation which occurs when deformation

ro
process is interrupted and the sample is held at a specific strain for EBSD data collection (those

troughs do not affect the microstructure.).


-p
re
lP
na
ur
Jo

10
of
ro
-p
re
lP
na
ur
Jo

Fig. 3 (a) An illustration of the EBSD test conducted on an in-situ deformed Ti2448 alloy sample

in which the directions of the X-, Y- and Z- axes are defined with respect to the sample; (b) the

11
stress-strain curve of the sample obtained from the in-situ EBSD test and inset images showing

the change of phase compositions of the sample in different deformation stages; (c) a plot showing

that the fraction of the α" phase increases with the increase of deformation strain; (d) (e) (f) (g) (h)

SEM images, (i) (j) (k) (l) (m) corresponding EBSD inverse polar figures and (n) (o) (p) (q) (r)

EBSD phase maps of the sample at different deformation strains of (d) (i) (n) 0%, (e) (j) (o) 1.2%,

(f) (k) (p) 3.3%, (g) (l) (q) 5% and (h) (m) (r) 7%.

of
Fig. 3 shows the SEM secondary electron images (Fig. 3(d)-(h)), EBSD inverse polar figures (Figs.

ro
3(i)-(m)) and EBSD phase maps (Figs. 3(n)-(r)) collected from the sample when it is in-situ

-p
deformed, with tensile strain varying from 0% to 7%. The black area in Fig. 3(d)-(h) and the white
re
area in Fig. 3(n)-(r) represent pores in the material. Grains of β phase can be seen clearly in Fig.
lP

3(d)-(r). Slip bands can be seen in grains in the sample deformed at strains of 3.3%, 5% and 7%.

Figs. 3(n)-(r) reveal the evolution of phase composition of the alloy during deformation. Both β
na

phase and α" phase can be identified by EBSD. They are represented by green and red regions
ur

respectively. Even in the undeformed sample (Fig. 3(n)), α" phase has been observed. This
Jo

observation contradicts the XRD result of the same sample which revealing a structure of pure β

phase. One possible explanation for contradiction is that martensite transformation has already

been triggered by the internal stress, which is introduced by rapid cooling when the alloy is water-

quenched from high temperature. However, the amount of α" phase is so little that it cannot be

distinguished by XRD. When comparing the amount of α" phase under different strains (Fig. 3(n)-

(r)), it is apparent that the fraction of the α" phase increases continuously as the deformation

process goes on. This trend is more clearly illustrated in the inset micrographs of a grain in Fig.

3(b), showing the change of phase compositions of the grain at different deformation stages. As

12
the deformation strain increases, the fraction of the β phase (in green color) decreases and is

transformed into α" phase (in red color).

In order to quantify the influence of martensite transformation on the deformation, fraction of the

red region (α" phase) in Figs. 3(n)-(r) are calculated using an image analysis software (Oxford

HKL Channel 5) afterwards. The increment of the fraction of the α" phase in the sample at each

deformation stage, ∆𝑓𝛼" , is defined as following:

of
𝜀 0
∆𝑓𝛼" = 𝑓𝛼" − 𝑓𝛼" (3)

ro
-p
where 𝑓𝛼"
𝜀
and 𝑓𝛼"
0
represent the fraction of α" phase in the sample deformed at a specific strain
re
and in the undeformed sample, respectively.
lP
na

Fig. 3(c) shows the increase of the fraction of α" phase in the sample under different tensile strains.

The results indicate that the fraction of α" phase remains unchanged during the elastic deformation
ur
Jo

stage. However, in the plastic deformation stage, it continuously increases as the tensile strain

increases. The result suggests that the martensite transformation only contributes to the plastic

deformation of the Ti2448 alloy, but has little influence on elastic deformation.

In subsequent analysis, we select an area which includes nearly 300 grains. The grains numbered

sequentially so that the fraction of the α" phase in each grain during deformation can be separately

recorded. The orientation preferences of grains prior to the tensile test and those having

experienced severe plastic deformation are investigated. The results shed light on the occurrence

conditions for martensite transformation. Fig. 4(a) shows the EBSD phase map of the whole area

13
of the Ti2448 alloy sample prior to the tensile test (at 0% deformation strain). Fig. 4(b) shows the

corresponding inverse polar figures obtained from all the grains of this area. The inverse polar

figures indicate that there is a texture structure existing in the sample with preferential orientations

〈001〉𝛽 //𝑋 − 𝑎𝑥𝑖𝑠 and 〈223〉𝛽 //𝑍 − 𝑎𝑥𝑖𝑠. In our experiment, the green samples are fabricated by

cold pressing in which the pressing force is applied in the Z- direction and lead to plastic

deformation happening in the powder. Therefore, it is believed that the texture structure found in

the sample before the tensile test is caused by the sample preparation method. In comparison, Fig.

of
4(c) shows the EBSD phase map of the selected area of the Ti2448 alloy sample with deformation

ro
strain of 7%. In order to find preferential orientation of the grains for martensite transformation,

-p
only grains with the fraction of the α" phase larger than 20% are picked up from the area and
re
shown in Fig. 4(c). Fig. 4(d) shows inverse polar figures obtained from those grains shown in Fig.
lP

4(c). Apparently, the orientation preference of the grains with a large fraction of α" phase is
na

different from that obtained from Fig. 4(a). Although the preferential orientation 〈223〉𝛽 //𝑍 − 𝑎𝑥𝑖𝑠

still exists in the grains, the preferential orientation 〈001〉𝛽//𝑋 − 𝑎𝑥𝑖𝑠 is no longer seen. Instead, a
ur
Jo

new preferential orientation, i.e., 〈113〉𝛽 //𝑋 − 𝑎𝑥𝑖𝑠, is identified from the grains with large fraction

of the α" phase. This result also agrees well with that reported by Yang et al.[8]. Their research

indicates that {332}〈113〉𝛽 twins play an important role in the formation of α" martensite in the

Ti2448 alloy.

14
of
ro
-p
re
Fig. 4 (a) EBSD phase map of the Ti2448 alloy sample at 0% deformation strain and (b)
lP

corresponding IPF maps indicating that a weak texture structure with preferential orientations
na

〈001〉𝛽 //𝑋 − 𝑎𝑥𝑖𝑠 𝑎𝑛𝑑 〈223〉𝛽 //𝑍 − 𝑎𝑥𝑖𝑠; (c) EBSD phase map of the grains with α" phase
ur

fraction larger than 20% selected from the sample at 7% deformation strain and (d) corresponding
Jo

IPF maps indicating preferential orientations different from the un-deformed sample, i.e.

〈113〉𝛽 //𝑋 − 𝑎𝑥𝑖𝑠 𝑎𝑛𝑑 〈223〉𝛽 //𝑍 − 𝑎𝑥𝑖𝑠.

To further determine the orientation relationships between β and α" phase, high-magnification

orientation color maps are shown in Fig. 5. Taking the highlighted β and α" phase as typical

examples, the α" grain distributes in the interior of β phase. The corresponding stereographic

15
projections shown in Fig. 5(b) indicate that α" grains share the same Burgers orientation

relationships with β phase. In other words, the α" grains tend to nucleate at either β grain interior

with specific orientation relationships, i.e., (11̅0)𝛽 //(110)𝛼" . This EBSD analyses are consistent

with TEM results (Fig. 2).

of
ro
-p
re
lP
na
ur
Jo

Fig. 5 High-magnification orientation color maps: (a) separate β and α" phase grains to investigate

their orientation relationships (strain: 5%); (b) stereographic projections of the β and α" grains

showing orientation relationships between β phase and α" phase.

5. Conclusions

In summary, Ti2448 alloy with good mechanical properties was fabricated by powder metallurgy

using element powders. The alloy exhibits good ductility in the tensile test. However, it has very

limited work hardening during plastic deformation. Its deformation mechanism is investigated by

in-situ EBSD test conducted in an SEM equipped with a dynamic tensile stage. The results indicate:

16
1. The strain hardening ratio of the Ti2448 alloy in this research is only 1.10, which is caused by

high oxygen content, leading to insignificant work hardening.

2. The amount of α" phase is so little that it cannot be distinguished by XRD. However, it is

apparent that the fraction of the α" phase increases continuously in the in-situ EBSD test as the

deformation process goes on.

3. The plastic deformation of the alloy has a dual model. Both dislocation slide and stress-induced

phase transformation contributes to the plastic deformation of the alloy. Deformation induced β

of
to α" martensitic transformation dominates the plastic deformation and strain hardening stages.

ro
4. Subsequent to yielding, β phase grains with preferential orientation <113> parallel to the tensile

-p
direction tend to be transformed into α" martensite phase, which reduces the deformation caused
re
by dislocation slide and accounts for the low strain hardening ratio of this alloy.
lP
na
ur

Data availability
Jo

The raw/processed data required to reproduce these findings cannot be shared at this time as the

data also forms part of an ongoing study.

CRediT authorship contribution statement

Xia Li: Conceptualization; Investigation; Roles/Writing - original draft; Writing - review &

editing. Jinhui Wang: Conceptualization; Data curation. Shulong Ye: Formal analysis; Funding

acquisition. Yinghao Zhou: Data curation; Methodology. Peng Yu: Project administration;

Supervision.

17
Declaration of competing interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This work is supported by the Shenzhen Science and Technology Innovation Committee (grant

number KQJSCX20180322152424539) and the China Postdoctoral Science Foundation (grant

of
number 2021M691401).

ro
-p
re
lP
na
ur
Jo

18
References

[1] S.L. Sing, F.E. Wiria, W.Y. Yeong, Selective laser melting of lattice structures: A statistical

approach to manufacturability and mechanical behavior, Robotics and Computer-Integrated

Manufacturing 49 (2018) 170-180.

[2] S. Ehtemam-Haghighi, H. Attar, M.S. Dargusch, D. Kent, Microstructure, phase composition

and mechanical properties of new, low cost Ti-Mn-Nb alloys for biomedical applications, Journal

of Alloys and Compounds 787 (2019) 570-577.

of
[3] Y. Yang, P. Castany, M. Cornen, F. Prima, S.J. Li, Y.L. Hao, T. Gloriant, Characterization of

ro
the martensitic transformation in the superelastic Ti–24Nb–4Zr–8Sn alloy by in situ synchrotron

-p
X-ray diffraction and dynamic mechanical analysis, Acta Materialia 88 (2015) 25-33.
re
[4] P. Castany, A. Ramarolahy, F. Prima, P. Laheurte, C. Curfs, T. Gloriant, In situ synchrotron
lP

X-ray diffraction study of the martensitic transformation in superelastic Ti-24Nb-0.5N and Ti-
na

24Nb-0.5O alloys, Acta Materialia 88 (2015) 102-111.

[5] T. Yao, K. Du, H. Wang, Z. Huang, C. Li, L. Li, Y. Hao, R. Yang, H. Ye, In situ scanning and
ur

transmission electron microscopy investigation on plastic deformation in a metastable β titanium


Jo

alloy, Acta Materialia 133 (2017) 21-29.

[6] X.G. Wang, L. Wang, M.X. Huang, Kinematic and thermal characteristics of Lüders and

Portevin-Le Châtelier bands in a medium Mn transformation-induced plasticity steel, Acta

Materialia 124 (2017) 17-29.

[7] Y.L. Hao, S.J. Li, S.Y. Sun, C.Y. Zheng, R. Yang, Elastic deformation behaviour of Ti–24Nb–

4Zr–7.9Sn for biomedical applications, Acta Biomaterialia 3(2) (2007) 277-286.

19
[8] Y. Yang, P. Castany, Y.L. Hao, T. Gloriant, Plastic deformation via hierarchical nano-sized

martensitic twinning in the metastable β Ti-24Nb-4Zr-8Sn alloy, Acta Materialia 194 (2020) 27-

39.

[9] B. Fer, D. Tingaud, A. Hocini, Y. Hao, E. Leroy, F. Prima, G. Dirras, Powder Metallurgy

Processing and Mechanical Properties of Controlled Ti-24Nb-4Zr-8Sn Heterogeneous

Microstructures, Metals 10(12) (2020) 1626.

[10] L.C. Zhang, D. Klemm, J. Eckert, Y.L. Hao, T.B. Sercombe, Manufacture by selective laser

of
melting and mechanical behavior of a biomedical Ti–24Nb–4Zr–8Sn alloy, Scripta Materialia 65(1)

ro
(2011) 21-24.

-p
[11] Y.J. Liu, X.P. Li, L.C. Zhang, T.B. Sercombe, Processing and properties of topologically
re
optimised biomedical Ti–24Nb–4Zr–8Sn scaffolds manufactured by selective laser melting,
lP

Materials Science and Engineering: A 642 (2015) 268-278.


na

[12] L. Zhang, T. Sercombe, Selective Laser Melting of Low-Modulus Biomedical Ti-24Nb-4Zr-

8Sn Alloy: Effect of Laser Point Distance, Key Engineering Materials 520 (2012) 226-233.
ur

[13] F. Kafkas, T. Ebel, Metallurgical and mechanical properties of Ti–24Nb–4Zr–8Sn alloy


Jo

fabricated by metal injection molding, Journal of Alloys and Compounds 617 (2014) 359-366.

[14] Y.J. Liu, H.L. Wang, S.J. Li, S.G. Wang, W.J. Wang, W.T. Hou, Y.L. Hao, R. Yang, L.C.

Zhang, Compressive and fatigue behavior of beta-type titanium porous structures fabricated by

electron beam melting, Acta Materialia 126 (2017) 58-66.

[15] J. Hernandez, S.J. Li, E. Martinez, L.E. Murr, X.M. Pan, K.N. Amato, X.Y. Cheng, F. Yang,

C.A. Terrazas, S.M. Gaytan, Y.L. Hao, R. Yang, F. Medina, R.B. Wicker, Microstructures and

Hardness Properties for β-Phase Ti–24Nb–4Zr–7.9Sn Alloy Fabricated by Electron Beam Melting,

Journal of Materials Science & Technology 29(11) (2013) 1011-1017.

20
[16] K.C. Nune, R.D.K. Misra, S.J. Li, Y.L. Hao, R. Yang, Cellular response of osteoblasts to low

modulus Ti-24Nb-4Zr-8Sn alloy mesh structure, Journal of Biomedical Materials Research Part A

105(3) (2017) 859-870.

[17] W. Kim, H. Kim, S. Lim, Effects of Oxygen on Phase Stability and Mechanical Properties of

Quenched Ti-Nb Alloys, Solid State Phenomena 124-126 (2007) 1377-1380.

[18] J. Niu, D. Ping, T. Ohno, W.-T. Geng, Suppression effect of oxygen on the β to ω

transformation in a β-type Ti alloy: Insights from first-principles, Modelling Simul. Mater. Sci.

of
Eng. 22 (2014) 5007.

ro
[19] H.-p. Duan, H.-x. Xu, W.-h. Su, Y.-b. Ke, Z.-q. Liu, H.-h. Song, Effect of oxygen on the

-p
microstructure and mechanical properties of Ti-23Nb-0.7Ta-2Zr alloy, International Journal of
re
Minerals, Metallurgy, and Materials 19(12) (2012) 1128-1133.
lP

[20] M. Besse, P. Castany, T. Gloriant, Mechanisms of deformation in gum metal TNTZ-O and
na

TNTZ titanium alloys: A comparative study on the oxygen influence, Acta Materialia 59(15) (2011)

5982-5988.
ur

[21] E.G. Obbard, Y.L. Hao, R.J. Talling, S.J. Li, Y.W. Zhang, D. Dye, R. Yang, The effect of
Jo

oxygen on α″ martensite and superelasticity in Ti–24Nb–4Zr–8Sn, Acta Materialia 59(1) (2011)

112-125.

[22] J.H. Dai, Y. Song, W. Li, R. Yang, L. Vitos, Influence of alloying elements Nb, Zr, Sn, and

oxygen on structural stability and elastic properties of the Ti2448 alloy, Physical Review B 89(1)

(2014) 014103.

[23] X. Li, S. Ye, X. Yuan, P. Yu, Fabrication of biomedical Ti-24Nb-4Zr-8Sn alloy with high

strength and low elastic modulus by powder metallurgy, Journal of Alloys and Compounds 772

(2019) 968-977.

21
[24] X. Li, Y. Zhou, T. Ebel, L. Liu, X. Shen, P. Yu, The influence of heat treatment processing

on microstructure and mechanical properties of Ti–24Nb–4Zr–8Sn alloy by powder metallurgy,

Materialia 13 (2020) 100803.

[25] M. Niinomi, Mechanical properties of biomedical titanium alloys, Materials Science and

Engineering: A 243(1) (1998) 231-236.

[26] M. Niinomi, Fatigue performance and cyto-toxicity of low rigidity titanium alloy, Ti-29Nb-

13Ta-4.6Zr, Biomaterials 24(16) (2003) 2673-2683.

of
[27] H.J. Rack, J.I. Qazi, Titanium alloys for biomedical applications, Materials ence &

ro
Engineering C 26(8) (2006) 1269-1277.

-p
[28] S. Bauer, P. Schmuki, K. von der Mark, J. Park, Engineering biocompatible implant surfaces:
re
Part I: Materials and surfaces, Progress in Materials Science 58(3) (2013) 261-326.
lP
na
ur
Jo

22
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

You might also like