You are on page 1of 8

Materials and Design 111 (2016) 592–599

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

Evaluation of mechanical and wear properties of Ti\\xNb\\7Fe alloys


designed for biomedical applications
S. Ehtemam-Haghighi a,⁎, K.G. Prashanth b, H. Attar a, A.K. Chaubey c, G.H. Cao d, L.C. Zhang a,⁎
a
School of Engineering, Edith Cowan University, 270 Joondalup Drive, Joondalup, Perth, WA 6027, Australia
b
Erich Schmid Institute of Materials Science, Austrian Academy of Sciences, Jahnstraße 12, A-8700 Leoben, Austria
c
Institute of Minerals and Materials Technology (IMMT), Bhubaneshwar, 751013, India
d
Department of Materials Engineering, Shanghai Key Laboratory of Modern Metallurgy and Materials Processing, Shanghai University, 149 Yanchang Road, Shanghai 200072, China

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Nb addition reduces the α" martensitic


phase content in the alloys microstruc-
ture.
• Nb content is inversely proportional to
the elastic modulus, hardness and com-
pressive strength of the alloys.
• The close relationship between the
wear resistance, microstructures and
mechanical properties of the alloys is
discussed.
• β-type Ti-11Nb-7Fe exhibits better
combination of properties than CP-Ti
and Ti-6Al-4V for orthopaedic applica-
tion.

a r t i c l e i n f o a b s t r a c t

Article history: A group of Ti\\xNb\\7Fe (x = 0, 1, 4, 6, 9, 11 wt.%) alloys was designed and produced by cold crucible levitation
Received 27 April 2016 melting process. The microstructural characteristic of the alloys with Nb addition and its effect on their mechan-
Received in revised form 6 September 2016 ical properties as well as wear resistance were investigated. Microscopic and phase analysis results show that all
Accepted 8 September 2016
the alloys, except for the Ti\\11Nb\\7Fe, exhibit orthorhombic α“ and body-centred cubic β phases, while
Available online 13 September 2016
Ti\\11Nb\\7Fe alloy consists of only β phase. It is proposed that increasing the Nb content enhances β phase sta-
Keywords:
bility and its proportion in the microstructure of the designed alloys. Depending on the proportion of β and α”
Titanium alloy phases, Ti\\xNb\\7Fe alloys show varied hardness (3.57–5.92 GPa) and compressive strength (1990–
Microstructure 2093 MPa). Additionally, they present wear rates in the range of 3 × 10−15–1 × 10−13 m3/m which correlates
Mechanical property well with the changes in the corresponding microstructures and mechanical properties. Among the studied al-
Wear loys, Ti\\11Nb\\7Fe with β phase microstructure, presents the lowest elastic modulus (86 GPa) and the highest
compressive strain (41.5%) along with high compressive strength, hardness and wear resistance. Therefore, it is
suggested that this β-type Ti\\11Nb\\7Fe alloy is a promising candidate, more suitable than the commercially
used CP\\Ti and Ti\\6Al\\4V, for orthopedic applications.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction
⁎ Corresponding authors.
E-mail addresses: sehtemam@our.ecu.edu.au, sehtemam@gmail.com An increasing aging population along with expectations of a high
(S. Ehtemam-Haghighi), l.zhang@ecu.edu.au, lczhangimr@gmail.com (L.C. Zhang). quality of life has led to a growing demand for the development of

http://dx.doi.org/10.1016/j.matdes.2016.09.029
0264-1275/© 2016 Elsevier Ltd. All rights reserved.
S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599 593

novel metallic biomaterials for replacing dysfunctional load-bearing work, the hardness of the alloys decreases with Fe addition up to
bone tissues [1]. Among metallic biomaterials, titanium (Ti) and its 7 mol% and subsequently increases by further Fe addition due to solid
alloys have been widely used for orthopedic applications due to solution strengthening. Their results showed that among the alloys, so-
their remarkable combination of good mechanical properties, high lution treated Ti\\5Fe\\3Nb\\3Zr possesses the lowest elastic modulus.
biocompatibility and excellent corrosion resistance [2–6]. However, Additionally, it shows higher tensile strength and better biocompatibil-
despite their attractive performance there are still challenges facing ity than the annealed Ti\\6Al\\4V ELI alloy [40].
their application in artificial joints [7]. These concerns are mainly re- Since there is relatively limited literature on the properties of Ti\\Fe
lated to bio-toxicity, the need for lower elastic modulus and better based alloys with Nb addition, this study intends to investigate the ef-
wear resistance. For instance, both Al and V elements in Ti\\6Al\\4V, fect of Nb on the microstructure of the designed Ti\\xNb\\7Fe alloys.
the most widely applied Ti alloy, are considered to induce long-term It also investigates its influence on the mechanical properties including
health problems such as haematological alterations and Alzheimer's hardness, elastic modulus, compressive yield strength, compressive
diseases [8,9]. Additionally, although the elastic modulus of Ti alloys strength and strain as well as wear responses of the alloys in order to as-
is lower than other common metallic biomaterials (e.g. 316 L stain- sess their suitability for orthopedic application. This study indicates that
less steel and Co\\Cr alloys), it is still higher than that of human the alloy design plays a key role in tailoring the mechanical and wear
bone [10,11]. The significant difference between the elastic modulus properties of Ti alloys required for biomedical application.
of the implant material and the surrounding bone may induce stress
shielding effect, whereby the bone is prevented from supporting 2. Experimental
loads useful to its health [12–14]. This may result in bone resorption
around the implant and eventual loosening of the implant [15–17]. A series of Ti\\xNb\\7Fe (x = 0, 1, 4, 6, 9, 11 wt.%) alloys was de-
Poor tribological properties have also been considered as a reason signed and fabricated from 99.9 wt.% pure Ti, Fe and Nb using a cold cru-
for the failure of an implant due to the release and accumulation of cible levitation melting method (CCLM) under the protection of an
metal wear debris in surrounding tissues, thereby causing adverse argon atmosphere in a water-cooled copper crucible. The CCLM furnace,
allergic tissue reactions and a reduction in the service life of the im- a type of induction melting furnace, was comprised of induction coils
plant [10]. Wear of dentures and wear of screws and plates used in which were wrapped around a water cooled copper crucible and con-
bone fracture repair are good examples of biomedical applications nected to a frequency inverter power supply. After the raw metals
where good tribological properties are needed [18]. Therefore, were completely melted and mixed adequately, the electric power
good wear properties play a key role in controlling and determining was cut off and the molten metals were quenched in the water-cooled
the long-term service of Ti alloys. copper crucible. The alloy ingots were flipped and re-melted four
To improve the wear resistance of Ti alloys a variety of methods have times to obtain homogeneity in their chemical composition. The dimen-
been explored, such as thin film surface modification [19], titanium ma- sions of as-cast alloys were 8 mm in diameter and 100 mm in length.
trix composites [20] and additive manufacturing technology for fabri- The as-cast samples were sectioned perpendicular to their longitudinal
cating Ti based materials [21]. However, weak adhesion of thin film to direction using a Buehler Isomet low speed diamond saw to obtain
substrate, undesirable increase in elastic modulus and poor surface specimens. Prior to microstructural study, the samples were ground
quality, respectively, have been the major constraints identified in and polished using standard metallographic procedures, followed by
using these methods [19,22,23]. etching in Kroll's solution (5 vol% HF, 30 vol% HNO3 and 65 vol% H2O).
Therefore, the design of new Ti alloys composed of non-toxic ele- These studies were then conducted using a scanning electron microsco-
ments showing not only high strength and hardness but also lower elas- py (SEM, JEOL JSM 6000). Constituent phases of the alloys were deter-
tic modulus and improved wear resistance when compared to the mined by X-ray diffraction (XRD) using a PANalytical EMPYREAN
conventional Ti alloys is desirable. It has been suggested that β-type Ti diffractometer with Co Kα radiation (λ = 0.1789 nm) operated at
alloys possess lower elastic modulus, higher strength and hardness as 40 kV and 40 mA. The integrated areas for α″ and β diffraction peaks
well as excellent abrasive and wear resistance than α and α + β-type in the XRD patterns were calculated by using the peak-fitting program,
Ti alloys [24–26]. Consequently, during the last decade, extensive re- Fityk, with Pearson VII function [41–43]. Based on the integrated areas,
search has focused on designing new β-type Ti alloys containing non- the volume fraction (Vf) of each phase was estimated in a similar man-
toxic elements such as Mo, Nb, Fe, Sn, Zr and Ta targeting the optimiza- ner to Ref [44] using the following equations:
tion of mechanical properties [27,28]. Examples of these alloys include
Ti\\Mo [29], Ti\\29Nb\\13Ta\\4.6Zr [30], Ti\\24Nb\\4Zr\\7.9Sn [31] Aαʺ
and Ti\\10Zr\\15Nb\\15Si [32]. V f ðαʺ Þ ¼ ð1Þ
Aαʺ þ Aβ
It has been reported that Fe is a non-toxic, low cost and biocompat-
ible β-stabilizer element for developing Ti alloys with high strength

[33–35], while biocompatible, β-stabilizer Nb contributes to a decrease V f ðβ Þ ¼ ð2Þ
Aαʺ þ Aβ
in the elastic modulus of the alloy, also required for orthopedic applica-
tion [36,37]. Hsu et al. [38] investigated the microstructure, bending
modulus, bending strength as well as biocompatibility of the as-cast where Vf(α″) and Vf(β) are the volume fractions and Aα″ and Aβ are the
Ti\\5Nb\\xFe alloys up to 5 wt.% Fe addition. Based on their study, total integrated areas, corresponding to α″ and β phases, respectively.
Ti\\5Nb\\2Fe exhibits the lowest bending modulus which is lower Evaluation of the mechanical properties of the alloys was based on
than that of CP\\Ti by 20%. Additionally, it shows the highest bending nanoindentation and compressive tests. The hardness and elastic mod-
strength to modulus ratio which is larger than that of CP\\Ti by 214%. ulus of the samples were measured by a nanoindentation test using a
Moreover, preliminary cell culturing results suggested that the UMIS equipment from Frischer-Cripps laboratories with a Berkovich in-
Ti\\5Nb\\xFe alloys are biocompatible and support cell attachment denter applying a maximum load of 400 mN. At least 25 indentations
[38]. Chaves et al. [39] studied the influence of phase transformation were carried out on the metallographically polished surface of each
on the elastic modulus of Ti\\xNb\\3Fe alloy system with Nb content sample and the average values of the hardness and elastic modulus
ranging from 10 to 25 wt.%. According to their study, increasing the were attained. To determine the compressive yield strength, compres-
Nb content improves the β phase stability and reduces the elastic mod- sive strength and strain of the alloys, a compressive test was carried
ulus of the alloys [39]. Ushida et al. [40] studied the mechanical proper- out on the cylindrical samples with a diameter of 8 mm and a height
ties and biocompatibility of Ti\\Fe\\Nb\\Zr alloy system with Fe of 16 mm using an Instron 5980 universal testing machine at a cross-
addition through thermo-mechanical treatments. According to their head speed of 0.001 mm/s.
594 S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599

Sliding wear tests were performed according to the ASTM G 99-05 alloys microstructure contains only the β phase. In order to study
standard at room temperature and in ambient atmospheric conditions whether ω phase is present in the studied alloys or not, XRD at lower
using a pin-on-disc device. A hard-faced stainless steel disc of 45 mm scanning speed was also carried out. Generally, the presence of ω
in diameter and 13 mm in thickness was applied against the as-cast al- phase could be detected by conducting XRD at lower scanning speed
loys (flat pins of 8 mm in diameter and 25 mm in height). The wear tests (0.5°/min) [1,45]. In the present study, the XRD patterns at lower scan-
were performed at a fixed load of 25 N with a sliding speed of 0.5 m/s for ning speed (not shown here) did not contain the peaks related to ω
15 min. The rate of wear was measured using the following equation phase suggesting the absence of this phase in the Ti\\xNb\\7Fe alloys
[21]: microstructure.
The SEM microstructures of the as-cast Ti\\7Fe and ternary
Vs Ti\\xNb\\7Fe alloys are shown in Fig. 2. Coexistence of the α″ martens-
Qs ¼ ð3Þ
Ls ite and β phase is evident in all alloys except for Ti\\11Nb\\7Fe in
which only the β phase is observed. As shown in Fig. 2(a), Ti\\7Fe ex-
where Qs is the wear rate, Vs is the sliding volume loss and Ls is the hibits the highest concentration of the α″ martensitic phase in the β ma-
sliding distance. The volume loss was measured from the wear loss de- trix. With the addition of Nb up to 9 wt.%, the proportion of the α″ phase
termined by calculating the weight of the flat head pins before and after decreases (Fig. 2(b–e)). With a further increase in Nb content to 11 wt.%,
tests. The sliding distance is given by Ls = 2π rs vs ts where rs is the radius the stability of the β phase is effectively enhanced so that the micro-
of the wear track (22.5 mm), vs is the speed in rounds per minutes (240 structure of the alloy consists entirely of β phase (Fig. 2(f)). It should
rpm) and ts is the time (15 min). Thereafter, the worn surface of the be noted that the microstructural observations are in good agreement
samples was evaluated using SEM. with the corresponding XRD patterns. The average β grain size of the
Ti\\xNb\\7Fe alloys was estimated to be in the range of 144–175 μm.
3. Results and discussions The average β grain size of the alloys decreases with increasing the Nb
contents, possibly due to Nb-grain boundary interaction which slows
Fig. 1 presents the XRD profiles of a series of Ti\\xNb\\7Fe alloys. the grain boundary growing [46].
Ti\\7Fe is comprised of the body-centered cubic (bcc) β phase with Based on the above results it can be deduced that the addition of
the least volume fraction (Vf, β = 62%) and the highest concentration more Nb to Ti\\7Fe alloy promotes the stabilization of the β phase
of the orthorhombic α″ martensitic phase (Vf, α″ = 38%). This fast and decreases the possibility of β to α″ martensitic transformation in
cooling induced α″ orthorhombic structure is obtained from a distorted the alloys during rapid cooling. To be more precise, by increasing the
hexagonal unit cell in which the c-axis of the orthorhombic unit cell cor- Nb content of the alloys, the β to α″ martensitic transformation temper-
responds to the c-axis of the hexagonal unit cell, whereas a and b corre- ature, Ms, decreases, hence the concentration of the α″ phase in the
late with the orthogonal axis of the hexagonal unit cell [4]. quenched alloys declines.
When 1 wt.% Nb is added, the resultant alloy shows the same crystal- Table 1 presents the hardness and elastic modulus values of the
line structure as that of Ti\\7Fe. However, the intensity of β phase peaks studied alloys measured by nanoindentation test. It can be inferred
increases, suggesting an increase in the volume fraction of the β phase from Table 1 that the aforementioned properties of the alloys are sensi-
(Vf, β = 66%) and reduction of that of the α″ phase (Vf, αʺ = 34%). tive to their β-stabilizer Nb content. The highest values for hardness
With addition of 4 or 6 wt.% Nb to Ti\\7Fe, the intensity of β peaks, (5.92 GPa) and elastic modulus (130 GPa) are observed in Ti\\7Fe
hence the proportion of the β phase increases (Vf, β = 82% and 90% re- alloy which has the highest proportion of α″ martensite in its micro-
spectively), while the number of the α″ phase peaks, and correspond- structure (Vf, α″ = 38%). By increasing the Nb content of the alloys up
ingly α″ phase concentration, decrease further. When Nb content is to 11 wt.%, hence decreasing the volume fraction of the α″ phase, both
9 wt.%, the alloy shows a very weak peak of the α″ phase in the XRD pat- hardness and elastic modulus of the alloys decline, reaching the mini-
tern suggesting that an increased amount of the β phase is retained in mum values of 3.57 GPa and 86 GPa, respectively, in the Ti\\11Nb\\7Fe
the alloy microstructure (Vf, β = 96%). Finally, with addition of alloy with an exclusively β phase microstructure. Consequently, the
11 wt.% Nb, the peaks corresponding to the α″ phase completely disap- progressive reduction in the measured hardness and elastic modulus
pear, while only peaks from the β phase are observed, indicating the values of the Ti\\xNb\\7Fe alloys with Nb addition can be ascribed to
the decrease in their α″ phase concentration. In fact, there are different
reports on elastic modulus and hardness of α″ and β phases in Ti alloys.
While Lee et al. [47] found that in their binary Ti\\Nb system the β and
α” phases exhibit similar values of hardness and elastic modulus,
Matlakhova et al. [48] and Zhu et al. [49] reported higher elastic modu-
lus and hardness for α″ than β phase which is consistent with the pres-
ent study. The hardness of all alloys examined in this study are larger
than that of the commercially used biomedical CP\\Ti (2.7 GPa) [32].
In addition, the hardness values for Ti\\xNb\\7Fe (x = 0, 1, 4 wt.%)
are greater than that of Ti\\6Al\\4V (4.2 GPa) [50]. Moreover, among
the studied alloys, Ti\\xNb\\7Fe (x = 6, 9, 11 wt.%) alloys, with higher
proportions of the β phase, present lower elastic modulus values than
those of CP\\Ti (112 GPa) [32] and Ti\\6Al\\4V (125 GPa) [50].
Table 1 also summarizes the compressive strength, 0.2% proof com-
pressive yield strength, and compressive strain of the Ti\\xNb\\7Fe al-
loys. Only Ti\\7Fe and Ti\\1Nb\\7Fe alloys failed during the test, while
other alloys deformed until the maximum load capacity (100 kN) of the
machine was reached and the test was stopped. According to Table 1,
Ti\\7Fe and Ti\\1Nb\\7Fe alloys, with the two highest α″ martensitic
phase concentrations, present the two lowest fracture strains (7.9%
and 10.0% respectively).
As the Nb content rises to 4 wt.%, the formation of the α″ phase dur-
Fig. 1. XRD profiles of as-cast Ti\
\xNb\
\7Fe alloys. ing quenching decreases, leading to a subsequent increase in the
S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599 595

Fig. 2. SEM microstructures of the etched Ti\


\xNb\
\7Fe alloys: (a) Ti\
\7Fe; (b) Ti\
\1Nb\
\7Fe; (c) Ti\
\4Nb\
\7Fe; (d) Ti\
\6Nb\
\7Fe; (e) Ti\
\9Nb\
\7Fe and (f) Ti\
\11Nb\
\7Fe.

maximum strain that the alloy endures. It is noted that further addition phase presents in the alloys microstructure as the Nb concentration in-
of Nb (6, 9 and 11 wt.%) results in enhancement of the compressive creases. In the case of Ti\\7Fe and Ti\\1Nb\\7Fe alloys, despite the
strain, reaching the highest value (41.5%) in Ti\\11Nb\\7Fe. Such in- presence of large concentrations of a high strength α″ phase (Fig. 2(a),
creases in the compressive strain with Nb addition can be attributed (b)), their compressive strength is lower than that of other
to the reduction in hardness of the alloys [51], which is caused by the Ti\\xNb\\7Fe alloys. This is because these two alloys are unable to en-
decrease in the proportion of the α″ phase present in their microstruc- dure more plastic deformation at high stresses due to the low the ductile
ture (Fig. 2). The high compressive strain indicates good workability of nature of the α″ phase, hence they fail in the early stages of plastic de-
the alloy at room temperature [52]. According to Table 1, except for formation. While other alloys, which contain lower concentrations of a
Ti\\7Fe and Ti\\1Nb\\7Fe, the compressive strength of the alloys de- low ductile α″ phase, do not fail during the test and present more plastic
creases with increasing Nb content. Since the orthorhombic crystal deformation. Therefore, their compressive strength and strain increase
structure of the α″ phase contains less slip systems than the bcc crystal compared to Ti\\7Fe and Ti\\1Nb\\7Fe alloys. It is noteworthy noting
structure of β phase, higher stress is required for plastic deformation of that the compressive strengths observed in the studied alloys are great-
the α″ phase than for the β phase matrix [52,53]. Consequently, the ob- er than those of typical biomedical Ti materials including CP\\Ti
served reduction tendency in the compressive strength of the (820 MPa) [54], Ti\\6Al\\4V (1300 MPa) [55] and even Ti\\TiB com-
Ti\\xNb\\7Fe (x = 4, 6, 9, 11 wt.%) alloys can be explained by taking posite materials (1420–1434 MPa) [23] produced by conventional and
into account the decrease in the volume fraction of a high strength α″ advanced manufacturing technologies.

Table 1
Hardness (H), elastic modulus (E), compressive strength (σmax), compressive yield strength (σy), and compressive strain (εmax) values obtained in Ti\
\xNb\
\7Fe alloys.

Alloy nominal composition (wt.%) H (GPa) E (GPa) σmax(MPa) σy(MPa) εmax(%)

Ti\
\7Fe 5.92 ± 0.06 130 ± 1 1990 ± 8 1847 ± 5 7.9 ± 0.2
Ti\
\1Nb\\7Fe 5.58 ± 0.09 124 ± 2 1972 ± 12 1785 ± 7 10.0 ± 0.9
Ti\
\4Nb\\7Fe 5.13 ± 0.04 115 ± 2 2093 ± 17 1539 ± 11 24.3 ± 0.3
Ti\
\6Nb\\7Fe 4.18 ± 0.09 97 ± 3 2014 ± 14 1144 ± 7 32.1 ± 1.7
Ti\
\9Nb\\7Fe 3.72 ± 0.04 89 ± 1 2007 ± 15 1010 ± 9 40.5 ± 1.9
Ti\
\11Nb\\7Fe 3.57 ± 0.03 86 ± 1 2006 ± 14 985 ± 8 41.5 ± 1.6
596 S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599

phase microstructure. The decreasing trend observed in the compres-


sive yield strength of the alloys can be attributed to reduction in the pro-
portion of the α″ phase which contains fewer slip systems than the β
phase, hence requires higher stresses for deformation as explained
above. It should be noted that the compressive yield strength of all stud-
ied Ti\\xNb\\7Fe alloys are higher than those of the commonly used
biomaterials, CP\\Ti (552 MPa) [4] and Ti\\6Al\\4V (970 MPa) [56].
The high yield strength enhances the capacity of the alloy against its
permanent shape change which could benefit patients [52].
Therefore, it can be concluded that all the Ti\\xNb\\7Fe alloys pres-
ent much higher compressive strengths and compressive yield
strengths, which potentially make them suitable materials for bearing
heavy loading in implant applications [2,8].
Classical theories of wear show that hardness is a key parameter in
controlling the wear of a material, and a hard material usually possess a
high wear resistance [57]. In fact, it is increasingly realized that the wear
of a material is related to its ability to resist elastic strain to failure, which
Fig. 3. H/E ratio and wear rate of the Ti\
\xNb\
\7Fe alloys. is determined by the H/E (hardness to elastic modulus) ratio [57,58]. A
higher H/E ratio indicates better wear resistance of a material [32]. Fig. 3
As can be inferred from Table 1, the compressive yield strengths of presents the H/E ratio and wear rate variations of the alloys with respect
the Ti\\xNb\\7Fe alloys are affected by Nb concentration. Among the to their Nb concentration. As can be seen, with the addition of Nb, the H/
studied alloys, Ti\\7Fe with the highest concentration of α″ phase pre- E of the alloys decreases. This trend indicates that the wear resistance of
sents the highest compressive yield strength. Conversely, the yield Ti\\xNb\\7Fe alloys reduces with Nb addition, which concurs with an in-
strength of the alloys reduces with increasing Nb concentration and crease in the alloys wear rate as demonstrated in Fig. 3. According to the
reaches a minimum in Ti\\11Nb\\7Fe alloy with an exclusively β results shown in this figure, Ti\\7Fe alloy, with the highest α″ martensite

Fig. 4. SEM micrographs showing the wear morphology of worn surfaces of Ti\
\xNb\
\7Fe alloys: (a) Ti\
\7Fe; (b) Ti\
\1Nb-7Fe; (c) Ti\
\4Nb\
\7Fe; (d) Ti\
\6Nb\
\7Fe; (e) Ti\
\9Nb\
\7Fe and (f)
Ti\
\11Nb\ \7Fe.
S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599 597

concentration, possesses the largest H/E ratio (0.0455), hence the lowest Additionally, the presence of ploughing grooves is observed. With a fur-
wear rate (3 × 10−15 m3/m). Upon addition of Nb, hence a decrease in ther increase in Nb content to 4 and 6 wt.%, the ploughing grooves get
the proportion of the α” phase (Fig. 2), the wear rate increases from deeper and delamination of the surface becomes more significant, pos-
1 × 10−14 m3/m for Ti\\1Nb\\7Fe to 1 × 10−13 m3/m for Ti\\11Nb\\7Fe sibly because of further reduction in the concentration of the α″ phase,
alloy with a single β phase microstructure. hence a decrease in the hardness of the alloys (Fig. 4(c–d)). Finally, with
As shown in Fig. 3, the Ti\\xNb\\7Fe alloys studied in this work pos- the subsequent addition of 9 and 11 wt.% Nb complete delamination of
sess a higher H/E ratio than the widely used Ti alloys, CP\\Ti (0.0241) some parts of the surface is observed which indicates higher material
[32] and Ti\\6Al\\4V (0.0336) [50]. It is therefore expected the alloys removal due to a further decrease in proportion of α″ phase, hence
in this study will present better wear resistance. For instance, the hardness of the alloys (Fig. 4(e–f)). It should be noted that the micro-
wear rate of cast CP\\Ti under similar conditions is observed to be structural observations of the samples after the wear test are consistent
1.55 × 10−12 m3/m [21]. with the wear rate trend shown in Fig. 3.
Fig. 4 presents the wear tracks of the cast alloys as a function of Nb Fig. 5 presents the XRD patterns and EDX spectra of the worn surface
content. The wear tracks of the alloys show the presence of typical in Ti\\7Fe and Ti\\11Nb\\7Fe alloys. The XRD results indicate that the
wear features including wear scars, ploughing grooves, microcracks worn surface of both alloys contain titanium oxide (TiO) (Fig. 5(a)). Ad-
and delamination. The Ti\\7Fe alloy shows shallow wear scars along ditionally, the XRD pattern of β-type Ti\\11Nb\\7Fe contains more TiO
with the presence of some delamination cracks (Fig. 4(a)), which are peaks than that of (α″ + β) Ti\\7Fe alloy, which indicates that more TiO
formed at the surface of the sample in contact with the counter disc is produced in the former alloy. The presence of a higher amount of ox-
due to high strain levels developed during the wear test [21]. However, ygen (9.31 wt.%) in Ti\\11Nb\\7Fe alloy than in Ti\\7Fe alloy
since this alloy possesses a high hardness (Table 1), the cracks are un- (2.24 wt.%) is also evident from the EDX spectra in Fig. 5(b).
able to propagate into a complete delamination of surface which indi- It has been reported that during the wear test, sliding of a specimen
cates a high resistance of this alloy to wear. As seen in Fig. 4(b), with against a counter disc causes the release of friction heat and a subse-
the addition of 1 wt.% of Nb to the Ti\\7Fe alloy, the wear surface of quent increase in the contact temperature of the worn surface [59].
the alloy shows deeper wear scars and fewer signs of delamination Due to the high reactivity of Ti, the high temperature can result in the
due to the lower hardness of this alloy compared to Ti\\7Fe. formation of oxides on the worn surface [60]. However, it has been
demonstrated that the oxides formed during wear are loosened and
brittle, hence do not provide protection from wear [59,61] and tend to
be continuously fragmented, producing wear debris [62,63]. Studies
have shown that according to Archard's wear equation [64], the wear
volume is inversely proportional to the hardness of the alloy under
investigation [64]. Therefore, in a lower hardness alloy, higher plas-
tic deformation of surface and sub-surface during wear causes
more delamination and production of a higher volume of metallic
wear debris which can be oxidized at high frictional temperatures
[59]. This leads to an increase in the roughness of the alloy surface
which results in a higher abrasion and wear rate [65] compared to
a higher hardness alloy, as can also be seen in Fig. 3. The existence
of more oxide particles in the lower hardness Ti\\11Nb\\7Fe alloy
is also evident in Fig. 4(f), in contrast with the higher hardness
Ti\\7Fe alloy shown in Fig. 4(a). As such, the higher oxygen content
in Ti\\11Nb\\7Fe alloy than in the Ti\\7Fe alloy (Fig. 5) can be ex-
plained by taking into account the lower hardness of the
Ti\\11Nb\\7Fe alloy which results in the presence of a higher pro-
portion of oxide debris on its worn surface.
The wear mechanisms observed in the Ti\\xNb\\7Fe samples are
similar to those of the widely used Ti alloys, CP\\Ti and Ti\\6Al\\4V,
where wear takes place as a result of abrasive and oxidative processes
[21,62]. The abrasive process is responsible for the ploughing grooves
and formation of cracks that propagate and lead to delamination of
the surface and material removal [21]. According to the results, although
the wear mechanisms operating in Ti\\xNb\\7Fe alloys, CP\\Ti and
Ti\\6Al\\4V are similar, the Ti\\xNb\\7Fe family of alloys are expected
to have a lower wear rate, or in other words, better wear resistance, be-
cause of the higher H/E ratio observed in the alloys which makes them
strong competitors for use in biomedical applications.
Based on these results, it can be concluded that addition of Nb con-
siderably affects the properties of the studied Ti\\xNb\\7Fe alloys.
Among these, Ti\\11Nb\\7Fe presents the best combination of me-
chanical properties, including the lowest elastic modulus (86 GPa) and
the highest compressive strain (41.5%). Additionally, it exhibits com-
pressive strength (2006 MPa) and compressive yield strength
(985 MPa) higher than those of CP\\Ti (820 MPa and 552 MPa) [4,54]
and Ti\\6Al\\4V (1300 MPa and 970 MPa) [55,56], as well as a hardness
(3.57 GPa) greater than that of CP\\Ti (2.7 GPa) [54]. Furthermore, de-
spite the higher wear rate of this alloy in comparison to the other stud-
Fig. 5. (a) XRD patterns and (b) EDX spectra of the Ti\
\7Fe and Ti\
\11Nb\
\7Fe alloys worn ied Ti\\xNb\\7Fe alloys, it is expected that it will possess better wear
surfaces. resistance compared to CP\\Ti and Ti\\6Al\\4V. The combination of
598 S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599

mechanical properties indicates that the Ti\\11Nb\\7Fe alloy has a manufactured by electron beam melting and selective laser melting, Acta Mater.
113 (2016) 56–67.
potential for use in orthopedic applications. [16] I. Dimić, I. Cvijović-Alagić, B. Völker, A. Hohenwarter, R. Pippan, Đ. Veljović, M. Rakin, B.
Bugarski, Microstructure and metallic ion release of pure titanium and Ti\ \13Nb\ \13Zr
4. Conclusion alloy processed by high pressure torsion, Mater. Des. 91 (2016) 340–347.
[17] Y.-H. Hon, J.-Y. Wang, Y.-N. Pan, Composition/phase structure and properties of tita-
nium-niobium alloys, Mater. Trans. 44 (2003) 2384–2390.
In this study, the effect of Nb on the microstructure of the designed [18] M.A. Hussein, A.S. Mohammed, N. Al-Aqeeli, Wear characteristics of metallic bioma-
Ti\\xNb\\7Fe alloys, the resulting mechanical properties and the terials: a review, Materials 8 (2015) 2749–2768.
[19] X. Liu, P.K. Chu, C. Ding, Surface modification of titanium, titanium alloys, and relat-
wear response of the alloys was investigated. The main results are sum-
ed materials for biomedical applications, Mater. Sci. Eng. R 47 (2004) 49–121.
marized as follows: [20] I. Kim, B. Choi, Y. Kim, Y. Lee, Friction and wear behavior of titanium matrix (TiB+
TiC) composites, Wear 271 (2011) 1962–1965.
1. The XRD analysis revealed that Ti\\7Fe alloy is comprised of β and α″ [21] H. Attar, K.G. Prashanth, A.K. Chaubey, M. Calin, L.C. Zhang, S. Scudino, J. Eckert,
phases. As more Nb is added to this alloy (1, 4, 6, 9 wt.%), the concen- Comparison of wear properties of commercially pure titanium prepared by selective
laser melting and casting processes, Mater. Lett. 142 (2015) 38–41.
tration of the α″ phase decreases, while the proportion of the β phase [22] E. Olakanmi, R. Cochrane, K. Dalgarno, A review on selective laser sintering/melting
increases reaching a maximum in Ti\\11Nb\\7Fe alloy with an ex- (SLS/SLM) of aluminium alloy powders: processing, microstructure, and properties,
clusively β phase microstructure. Prog. Mater. Sci. 74 (2015) 401–477.
[23] H. Attar, M. Bönisch, M. Calin, L.-C. Zhang, S. Scudino, J. Eckert, Selective laser melt-
2. The microstructural studies performed by XRD analysis and SEM ing of in situ titanium–titanium boride composites: processing, microstructure and
showed that increasing the Nb content enhances the stability of the mechanical properties, Acta Mater. 76 (2014) 13–22.
[24] Y. Hao, Z. Zhang, S. Li, R. Yang, Microstructure and mechanical behavior of a
designed alloys against β to α″ martensitic transformation during
Ti\ \24Nb\ \4Zr\ \8Sn alloy processed by warm swaging and warm rolling, Acta
quenching. Mater. 60 (2012) 2169–2177.
3. The compressive strength (1990–2093 MPa), compressive yield [25] L. Zou, Y. Li, C. Yang, S. Qu, Y. Li, Effect of Fe content on glass-forming ability and
strength (985–1847 MPa) and hardness (3.57–5.92 GPa) of the crystallization behavior of a (Ti 69.7 Nb 23.7 Zr 4.9 Ta 1.7) 100− xFex alloy synthe-
sized by mechanical alloying, J. Alloys Compd. 553 (2013) 40–47.
Ti\\xNb\\7Fe alloys are higher than those of the widely used CP\\Ti [26] M.A.-H. Gepreel, M. Niinomi, Biocompatibility of Ti-alloys for long-term implanta-
(820 MPa, 552 MPa and 2.7 GPa,) or Ti\\6Al\\4V (1300 MPa, tion, J. Mech. Behav. Biomed. Mater. 20 (2013) 407–415.
970 MPa and 4.2 GPa) materials. [27] D. Gordin, A. Guillou, I. Thibon, M. Bohn, D. Ansel, T. Gloriant, Duplex nitriding treat-
ment of a beta-metastable Ti 94 Mo 6 alloy for biomedical applications, J. Alloys
4. Among the studied alloys, Ti\\7Fe exhibits the highest elastic modu- Compd. 457 (2008) 384–388.
lus (130 GPa), while β-type Ti\\11Nb\\7Fe alloy shows the lowest [28] Y. Li, L. Zou, C. Yang, Y. Li, L. Li, Ultrafine-grained Ti-based composites with high
elastic modulus (86 GPa) which is smaller than those of CP\\Ti strength and low modulus fabricated by spark plasma sintering, Mater. Sci. Eng. A
560 (2013) 857–861.
(112 GPa) and Ti\\6Al\\4V (125 GPa). [29] Y.-y. Chen, L.-j. Xu, Z.-g. Liu, F.-t. Kong, Z.-y. Chen, Microstructures and properties of
5. The wear resistance of Ti\\xNb\\7Fe alloys reduces by increasing titanium alloys Ti\ \Mo for dental use, Trans. Nonferrous Metals Soc. China 16 (2006)
their Nb content. However, it is expected that alloys will show better s824–s828.
[30] M. Niinomi, Fatigue performance and cyto-toxicity of low rigidity titanium alloy,
wear resistance than the widely used CP-Ti and Ti\\6Al\\4V.
Ti\\29Nb\ \13Ta\ \4.6 Zr, Biomaterials 24 (2003) 2673–2683.
6. Among the alloys, Ti\\11Nb\\7Fe presents a favourable combination [31] Y. Hao, S. Li, S. Sun, C. Zheng, Q. Hu, R. Yang, Super-elastic titanium alloy with unsta-
of properties for orthopedic application. ble plastic deformation, Appl. Phys. Lett. 87 (2005) 91906.
[32] S. Abdi, M.S. Khoshkhoo, O. Shuleshova, M. Bönisch, M. Calin, L. Schultz, J. Eckert, M.
Baró, J. Sort, A. Gebert, Effect of Nb addition on microstructure evolution and nano-
References mechanical properties of a glass-forming Ti\ \Zr\
\Si alloy, Intermetallics 46 (2014)
156–163.
[1] H.-C. Hsu, S.-C. Wu, S.-K. Hsu, T.-F. Lin, W.-F. Ho, Structure and mechanical proper- [33] J. Lu, Y. Zhao, H. Niu, Y. Zhang, Y. Du, W. Zhang, W. Huo, Electrochemical corrosion
ties of as-cast Ti\ \5Nb\ \xCr alloys, Mater. Des. 51 (2013) 268–273. behavior and elasticity properties of Ti\ \6Al\ \xFe alloys for biomedical applications,
[2] P. Wang, Y. Feng, F. Liu, L. Wu, S. Guan, Microstructure and mechanical properties of Mater. Sci. Eng. C (2016).
Ti\\Zr\\Cr biomedical alloys, Mater. Sci. Eng. C 51 (2015) 148–152. [34] D. Kuroda, H. Kawasaki, A. Yamamoto, S. Hiromoto, T. Hanawa, Mechanical proper-
[3] S. Gabriel, J. Panaino, I. Santos, L. Araujo, P. Mei, L. de Almeida, C. Nunes, Character- ties and microstructures of new Ti\ \Fe\\Ta and Ti\ \Fe\\Ta\\Zr system alloys, Mater.
ization of a new beta titanium alloy, Ti\ \12Mo\ \3Nb, for biomedical applications, J. Sci. Eng. C 25 (2005) 312–320.
Alloys Compd. 536 (2012) S208–S210. [35] Y. Abd-elrhman, M.A.-H. Gepreel, A. Abdel-Moniem, S. Kobayashi, Compatibility as-
[4] S.E. Haghighi, H. Lu, G. Jian, G. Cao, D. Habibi, L.C. Zhang, Effect of α″martensite on sessment of new V-free low-cost Ti–4.7 Mo–4.5 Fe alloy for some biomedical appli-
the microstructure and mechanical properties of beta-type Ti\ \Fe\ \Ta alloys, cations, Mater. Des. 97 (2016) 445–453.
Mater. Des. 76 (2015) 47–54. [36] A. Nouri, P.D. Hodgson, C. Wen, Effect of ball-milling time on the structural charac-
[5] N. Dai, L.-C. Zhang, J. Zhang, Q. Chen, M. Wu, Corrosion behavior of selective laser teristics of biomedical porous Ti\ \Sn\\Nb alloy, Mater. Sci. Eng. C 31 (2011)
melted Ti\ \6Al\ \4V alloy in NaCl solution, Corros. Sci. 102 (2016) 484–489. 921–928.
[6] G. Purcek, G. Yapici, I. Karaman, H. Maier, Effect of commercial purity levels on the [37] I. Kopova, J. Stráský, P. Harcuba, M. Landa, M. Janeček, L. Bačákova, Newly developed
mechanical properties of ultrafine-grained titanium, Mater. Sci. Eng. A 528 (2011) Ti\\Nb\ \Zr\ \Ta\ \Si\
\Fe biomedical beta titanium alloys with increased strength and
2303–2308. enhanced biocompatibility, Mater. Sci. Eng. C 60 (2016) 230–238.
[7] S. Wang, Z. Ma, Z. Liao, J. Song, K. Yang, W. Liu, Study on improved tribological prop- [38] H.-C. Hsu, S.-K. Hsu, S.-C. Wu, C.-J. Lee, W.-F. Ho, Structure and mechanical proper-
erties by alloying copper to CP\ \Ti and Ti\ \6Al\ \4V alloy, Mater. Sci. Eng. C 57 ties of as-cast Ti\ \5Nb\ \xFe alloys, Mater. Charact. 61 (2010) 851–858.
(2015) 123–132. [39] J. Chaves, O. Florêncio, P. Silva, P. Marques, C. Afonso, Influence of phase transforma-
[8] Y. Liu, K. Li, T. Luo, M. Song, H. Wu, J. Xiao, Y. Tan, M. Cheng, B. Chen, X. Niu, Powder tions on dynamical elastic modulus and anelasticity of beta Ti\ \Nb\ \Fe alloys for
metallurgical low-modulus Ti\ \Mg alloys for biomedical applications, Mater. Sci. biomedical applications, J. Mech. Behav. Biomed. Mater. 46 (2015) 184–196.
Eng. C 56 (2015) 241–250. [40] K. Ushida, K. Tsuge, T. Akahori, T. Hattori, M. Niinomi, K. Ishikura, M.A. Gepreel, Me-
[9] M. Atapour, A. Pilchak, G. Frankel, J. Williams, Corrosion behavior of β titanium chanical strength and biocompatibility of meta-stable[beta] type Ti-5 Fe-3 Nb-3 Zr
alloys for biomedical applications, Mater. Sci. Eng. C 31 (2011) 885–891. for biomedical applications, J. Jpn. Inst. Met. 76 (2012) 397–401.
[10] M. Geetha, A. Singh, R. Asokamani, A. Gogia, Ti based biomaterials, the ultimate [41] Sh. Ehtemam Haghighi, K. Janghorban, S. Izadi, Order-sintering of mechanically
choice for orthopaedic implants—a review, Prog. Mater. Sci. 54 (2009) 397–425. alloyed FeAl nanostructures, J. Alloys Compd. 503 (2010) 375–379.
[11] L.-C. Zhang, H. Attar, Selective laser melting of titanium alloys and titanium matrix [42] Sh. E. Haghighi, K. Janghorban, S. Izadi, Structural evolution of Fe–50 at.% Al powders
composites for biomedical applications: a review, Adv. Eng. Mater. 18 (2016) during mechanical alloying and subsequent annealing processes, J. Alloys Compd.
463–475. 495 (2010) 260–264.
[12] A. Biesiekierski, J. Lin, Y. Li, D. Ping, Y. Yamabe-Mitarai, C. Wen, Investigations into [43] L.C. Zhang, Z. Shen, J. Xu, Glass formation in a (Ti, Zr, Hf)–(Cu, Ni, Ag)–Al high-order
Ti\\(Nb, Ta)\ \Fe alloys for biomedical applications, Acta Biomater. (2015). alloy system by mechanical alloying, J. Mater. Res. 18 (2003) 2141–2149.
[13] P.F. Santos, M. Niinomi, H. Liu, K. Cho, M. Nakai, Y. Itoh, T. Narushima, M. Ikeda, Fab- [44] H. Yang, J. Wen, M. Quan, J. Wang, Evaluation of the volume fraction of nanocrystals
rication of low-cost beta-type Ti\ \Mn alloys for biomedical applications by metal in- devitrified in Al-based amorphous alloys, J. Non-Cryst. Solids 355 (2009) 235–238.
jection molding process and their mechanical properties, J. Mech. Behav. Biomed. [45] W.-F. Ho, S.-C. Wu, H.-H. Chang, H.-C. Hsu, Structure and mechanical properties of
Mater. 59 (2016) 497–507. Ti\\5Cr based alloy with Mo addition, Mater. Sci. Eng. C 30 (2010) 904–909.
[14] I. Hacisalihoglu, A. Samancioglu, F. Yildiz, G. Purcek, A. Alsaran, Tribocorrosion prop- [46] L.-b. Zhang, K.-z. Wang, L.-j. Xu, S.-l. Xiao, Y.-y. Chen, Effect of Nb addition on micro-
erties of different type titanium alloys in simulated body fluid, Wear 332 (2015) structure, mechanical properties and castability of β-type Ti\ \Mo alloys, Trans. Non-
679–686. ferrous Metals Soc. China 25 (2015) 2214–2220.
[15] Y. Liu, S. Li, H. Wang, W. Hou, Y. Hao, R. Yang, T. Sercombe, L.C. Zhang, Microstruc- [47] C. Lee, C.-P. Ju, J. Chern Lin, Structure–property relationship of cast Ti\ \Nb alloys, J.
ture, defects and mechanical behavior of beta-type titanium porous structures Oral Rehabil. 29 (2002) 314–322.
S. Ehtemam-Haghighi et al. / Materials and Design 111 (2016) 592–599 599

[48] L. Matlakhova, A. Matlakhov, S. Monteiro, S. Fedotov, B. Goncharenko, Properties [58] A. Hynowska, E. Pellicer, J. Fornell, S. González, N. van Steenberge, S. Suriñach, A.
and structural characteristics of Ti\ \Nb\ \Al alloys, Mater. Sci. Eng. A 393 (2005) Gebert, M. Calin, J. Eckert, M.D. Baró, Nanostructured β-phase Ti–31.0 Fe–9.0 Sn
320–326. and sub-μm structured Ti–39.3 Nb–13.3 Zr–10.7 Ta alloys for biomedical applica-
[49] Y. Zhu, X. Wang, L. Wang, Y. Fu, J. Qin, W. Lu, D. Zhang, Influence of forging deforma- tions: microstructure benefits on the mechanical and corrosion performances,
tion and heat treatment on microstructure of Ti\ \xNb\ \3Zr\\2Ta alloys, Mater. Sci. Mater. Sci. Eng. C 32 (2012) 2418–2425.
Eng. C 32 (2012) 126–132. [59] X. Li, Y. Zhou, X. Ji, Y. Li, S. Wang, Effects of sliding velocity on tribo-oxides and wear
[50] W. Xue, C. Wang, R. Chen, Z. Deng, Structure and properties characterization of ce- behavior of Ti\ \6Al\ \4V alloy, Tribol. Int. 91 (2015) 228–234.
ramic coatings produced on Ti\ \6Al\ \4V alloy by microarc oxidation in aluminate [60] P. La, J. Ma, Y.T. Zhu, J. Yang, W. Liu, Q. Xue, R.Z. Valiev, Dry-sliding tribological prop-
solution, Mater. Lett. 52 (2002) 435–441. erties of ultrafine-grained Ti prepared by severe plastic deformation, Acta Mater. 53
[51] K. Cho, M. Niinomi, M. Nakai, J. Hieda, Y. Kawasaki, Development of high modulus (2005) 5167–5173.
Ti\\Fe\\Cu alloys for biomedical applications, Mater. Trans. 54 (2013) 574–581. [61] K. Farokhzadeh, A. Edrisy, Transition between mild and severe wear in titanium
[52] S. Ehtemam-Haghighi, Y. Liu, G. Cao, L.-C. Zhang, Phase transition, microstructural alloys, Tribol. Int. 94 (2016) 98–111.
evolution and mechanical properties of Ti\ \Nb\ \Fe alloys induced by Fe addition, [62] G. Straffelini, A. Molinari, Dry sliding wear of Ti\ \6Al\\4V alloy as influenced by the
Mater. Des. 97 (2016) 279–286. counterface and sliding conditions, Wear 236 (1999) 328–338.
[53] Y. Ren, F. Wang, S. Wang, C. Tan, X. Yu, J. Jiang, H. Cai, Mechanical response and ef- [63] A. Alsaran, G. Purcek, I. Hacisalihoglu, Y. Vangolu, Ö. Bayrak, I. Karaman, A. Celik, Hy-
fects of β-to-α″ phase transformation on the strengthening of Ti\ \10V\ \2Fe\\3Al droxyapatite production on ultrafine-grained pure titanium by micro-arc oxidation
during one-dimensional shock loading, Mater. Sci. Eng. A 562 (2013) 137–143. and hydrothermal treatment, Surf. Coat. Technol. 205 (2011) S537–S542.
[54] F.-W. Long, Q.-W. Jiang, L. Xiao, X.-W. Li, Compressive deformation behaviors of [64] Q. Wang, P.-Z. Zhang, D.-B. Wei, X.-H. Chen, R.-N. Wang, H.-Y. Wang, K.-T. Feng, Mi-
coarse-and ultrafine-grained pure titanium at different temperatures: a compara- crostructure and sliding wear behavior of pure titanium surface modified by double-
tive study, Mater. Trans. 52 (2011) 1617–1622. glow plasma surface alloying with Nb, Mater. Des. 52 (2013) 265–273.
[55] K. Srinivasan, P. Venugopal, Compression testing of Ti\ \6Al\\4V in the temperature [65] G. Purcek, O. Saray, O. Kul, I. Karaman, G. Yapici, M. Haouaoui, H. Maier, Mechanical
range of 303–873 K, Mater. Manuf. Process. 23 (2008) 342–346. and wear properties of ultrafine-grained pure Ti produced by multi-pass equal-
[56] W. Yan, J. Berthe, C. Wen, Numerical investigation of the effect of porous titanium channel angular extrusion, Mater. Sci. Eng. A 517 (2009) 97–104.
femoral prosthesis on bone remodeling, Mater. Des. 32 (2011) 1776–1782.
[57] J. Xu, G.d. Wang, X. Lu, L. Liu, P. Munroe, Z.-H. Xie, Mechanical and corrosion-
resistant properties of Ti\ \Nb\ \Si\ \N nanocomposite films prepared by a double
glow discharge plasma technique, Ceram. Int. 40 (2014) 8621–8630.

You might also like