You are on page 1of 7

Materials Science & Engineering A 760 (2019) 339–345

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Evaluation of the mechanical and wear properties of titanium produced by T


three different additive manufacturing methods for biomedical application
H. Attara,*, M.J. Berminghama, S. Ehtemam-Haghighia,b, A. Dehghan-Manshadia, D. Kenta,b,
M.S. Darguscha
a
Queensland Centre for Advanced Materials Processing and Manufacturing, School of Mechanical and Mining Engineering, The University of Queensland, Brisbane,
Queensland, 4072, Australia
b
School of Science and Engineering, University of the Sunshine Coast, Maroochydore DC, QLD, 4558, Australia

A R T I C LE I N FO A B S T R A C T

Keywords: Commercially pure titanium, as a widely used metallic biomaterial, was fabricated using dissimilar additive
Additive manufacturing manufacturing (AM) methods, namely selective laser melting (SLM), laser engineered net shaping (LENS) and
Titanium wire arc additive manufacturing (WAAM). Microstructures as well as mechanical and wear properties of the
Mechanical properties produced titanium samples were studied. Diverse microstructural features were related to the different linear
Wear resistance
energy densities and cooling rates induced by each AM method. Tensile testing evaluation indicated the highest
Microstructure
yield and ultimate tensile strengths as well as elastic energy for titanium produced by SLM. However, the
maximum ductility was obtained in the WAAM-fabricated titanium due to its larger grain size and slightly higher
densification. All the mechanical properties obtained were either superior or comparable to those of cast and
powder metallurgy produced titanium. Fracture surface analysis showed the presence of mainly coarse and fine
dimples for WAAM and SLM-produced samples, respectively. This was consistent with the grain size of each
sample. Wear performances and mechanisms were also examined and the results were in agreement with the
values obtained from the hardness to elastic modulus ratios (H/E and H3/E2).

1. Introduction various additive manufacturing (AM) methods as outlined in ISO/ASTM


52900. The metal AM methods have been described elsewhere and they
Ti and Ti alloys are widely used for various areas including bio- have generally substantial similarities [16]. However, some of them
medical applications due to their outstanding biocompatibility, corro- differ in important aspects such as the material feeding system and/or
sion resistance and mechanical properties [1–3]. Over the past years, a power source configuration which influence melting, solidification and,
large number of Ti alloys, especially alloys from the β group, have been consequently, microstructures as well as resulting performances, in-
developed for biomedical applications [4,5]. Although β-Ti alloys are cluding mechanical and wear properties. It is important that AM-pro-
increasingly being developed for orthopedic applications, they gen- duced parts show properties at least comparable or even better to those
erally contain alloying elements with densities and melting tempera- fabricated by traditional techniques to fulfill the standards required for
tures higher than commercially pure Ti (CP–Ti). This may increase the biomedical orthopedic applications.
density of the alloy, introducing additional processing complexities and Although AM of Ti materials has been progressively studied, far less
increase the final cost of the product [6,7]. Therefore, biomedical im- attention has been paid to comparative studies for a given Ti material.
plants fabricated from CP-Ti are still of interest, despite their relatively Over the past years, a number of comparative studies have been con-
lower mechanical properties compared to Ti alloys. ducted on Ti–6Al–4V alloy produced by different AM and conventional
Ti-based materials are generally fabricated by conventional tech- manufacturing methods [17,18], however, AM methods applied in the
niques such as casting [8] and powder metallurgy [9]. Over recent aforementioned studies have the same material feeding systems and
years, additive manufacturing has been also applied increasingly for the nearly similar power source configurations. To our best knowledge,
fabrication of Ti materials due to its advantages over conventional limited attention has been devoted to study the influence of dissimilar
manufacturing methods. These advantages mainly include the ability to AM technologies on the important properties of CP-Ti. Therefore, this
produce near net shapes and lower production times [10–15]. There are work aims to study the processing conditions as well as the

*
Corresponding author.
E-mail addresses: hooyar.attar@gmail.com, h.attar@uq.edu.au (H. Attar).

https://doi.org/10.1016/j.msea.2019.06.024
Received 2 November 2018; Received in revised form 9 May 2019; Accepted 7 June 2019
Available online 08 June 2019
0921-5093/ © 2019 Elsevier B.V. All rights reserved.
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

microstructures, mechanical and wear properties of CP-Ti produced by titanium samples was assessed based on the weight loss method by
different AM methods (including Powder Bed Fusion and Direct Energy measuring the weight of the samples before and after the wear test
Deposition specified by ASTM 52900), namely selective laser melting using an accurate microbalance. Additionally, a Hitachi (TM3030)
(SLM), laser engineered net shaping (LENS) and wire arc additive scanning electron microscope (SEM) was applied to study the worn
manufacturing (WAAM). These AM methods utilize different feedstock surfaces as well as the fracture surface of the specimens after tensile
material and feeding systems specifically powder-bed, blown-powder testing.
and wire-feeding, respectively, and different heat-input methods, em-
ploying laser melting for SLM and LENS, and an electric arc for WAAM. 3. Results and discussion
This will lead to varied interactions between the power sources and
starting materials as well as different scan speeds, energy densities, melt Fig. 1 shows the microstructures of the AM-produced titanium
pool sizes, heating/cooling rates and consequently process-micro- samples. The microstructure of the SLM-produced titanium reveals the
structure-properties relationships. This study will use linear energy martensitic phase (α′) with lath-type and acicular morphologies. In
density as one of the key factors that can be used to compare the dif- contrast, titanium fabricated using LENS and WAAM show a structure
ferent AM processes. The main focus of the paper is to provide a dominated by the α phase structure with various morphologies. The
technical framework which can guide researchers and practitioners in LENS-produced titanium displays mainly plate-like α, with serrated and
the AM and biomedical implant engineering communities to select the acicular α in some regions. Similarly, WAAM-produced titanium exhibit
most desirable manufacturing method according to their requirements mainly serrated α with limited amounts of acicular α in some areas. It is
and resources. also noted that SLM-produced titanium shows the finest grain size
(1.8 μm), while WAAM-fabricated titanium demonstrates the largest
2. Experimental grain size (11.9 μm) as presented in Table 2. Diverse microstructural
features can be explained through interrelated varied energy densities
Commercially pure titanium (CP–Ti) in the forms of both powder and cooling rates. It is known that the scale of the α-phase in titanium
and wire was used in this study. The powder and wire feedstock com- and its alloys is influenced by the cooling rates during its formation (i.e.
plied with the composition of CP-Ti (ASTM Grade 1). The feedstock during the β → α transition) [20]. As reported previously [21], the ratio
materials were supplied by Advanced Powders and Coatings Company of power to scanning speed is described by the linear energy density
and Baoji Xinnuo Net Metal Materials Company, respectively. The (LED), a key determinant of the heating environment:
chemical composition of both powder and wire feedstock are listed in P
Table 1. To fabricate the titanium samples, SLM Solution 250HL and LED =
V (1)
Optomec LENS™ 450 machines, and a WAAM deposition device with
the processing details explained previously [19] were used. The AM Where P is the heating power (W) and V is the scanning speed (mm/s).
processing parameters used to obtain the highest possible densifications The differing phases and microstructural features of the titanium sam-
are listed in Table 2. Although not the focus of the present study, the ples can be related to the different heating powers and scanning speeds
deposition parameters in Table 2 were determined from pilot studies to which impart varied LEDs (Table 2) induced by dissimilar AM pro-
identify suitable processing parameters required to produce titanium cesses. The power source for WAAM is much higher than those of LENS
samples with at least 99.5% densification measured by Archimedes and SLM, while the scanning speed applied in WAAM is considerably
method. Deposited samples were then cut and metallographic speci- slower. Hence, a significantly higher heat input was formed in WAAM
mens were prepared through standard procedures and they were also compared to LENS and especially to SLM. As energy density influences
etched using a solution consisting of 10% HF, 5% HNO3, and 85% the heat diffusion depth, the penetration distance will be much higher
distilled water (all in volume fractions). They were then examined using in WAAM than LENS and SLM. It is thus expected that when new ti-
a Polyvar Met optical microscope for microstructural characterizations. tanium layers are formed during WAAM on the previously deposited
To evaluate the mechanical properties, tensile and microhardness layers, heat removal through the preformed layers and substrate is slow
tests were conducted. Tensile specimens with same geometry and di- due to accumulated heat. This also led to some minor sample distortion
mensions (4 mm diameter and 20 mm gauge length as per ASTM E8M) in the WAAM-produced titanium induced by thermal residual stress
were prepared by machining from the AM-produced bulk samples. which was resolved through short stress-relief treatments at 480 °C for
Tensile testing was performed on the prepared round titanium speci- 2 h (recommended stress relief for CP-Ti that does not cause grain
mens using an Instron machine with a crosshead speed of 0.001 mm/s. growth [22]). The development of the observed microstructure can be
The Instron machine was equipped with extensometer for strain mea- also interpreted through an understanding of the solidification and
surement. Young's modulus, yield and ultimate strengths and de- cooling rates using the following equation [23]:
formation strain (ductility) of the tested specimens were measured from
1 ∂T
the stress–strain curves. Tensile tests were performed at least three R=
G ∂t (2)
times for each group. Vickers microhardness measurements were car-
ried out on the polished surface of the titanium samples using a Struers where R is the solidification rate, G is the thermal gradient and ∂T is the
∂t
Duramin hardness tester with a load of 50 g and 15 s dwell time. The cooling rate. In our study, WAAM-produced titanium under the afore-
average microhardness values were obtained from at least 10 mea- mentioned processing conditions has the slowest solidification and
surements for each sample. To evaluate the wear performance, a pin- cooling rates, leading to an increased α grain size in the phase trans-
on-disk device was used and the cylindrical titanium samples were formation of β to α. Similar conditions also occurred during LENS
abraded against a hardened steel disk under 30 N load, with 0.54 m/s which led to the formation of α phase with a relatively larger size
sliding speed, and 250 m sliding distance. Wear resistance of the compared to SLM-produced titanium but lower size than the WAAM

Table 1
Chemical compositions of the starting CP-Ti powder and wire feedstock used in the current work (wt.%).
Material type Supplier Ti C Fe O N

CP-Ti powder used for SLM and LENS Advanced Powders and Coatings (AP&C) Balance 0.02 0.02 0.07 0.02
CP-Ti wire used for WAAM Baoji Xinnuo Net Metal Materials Balance 0.009 0.023 0.05 0.014

340
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

Table 2
List of the processing parameters used for AM of titanium samples; the calculated LEDs and the resulting densifications and grain widths.
Method Power (W) Scanning speed (mm/s) LED (J/mm) Densification (%) Grain width (μm)

SLM 165 138 1.19 99.6 ± 0.1 1.8 ± 0.9


LENS 350 12 29.16 99.5 ± 0.2 3.1 ± 1.1
WAAM 1237 0.83 1490 99.7 ± 0.1 11.9 ± 2.6

Fig. 1. Optical microstructural images of CP-Ti samples produced by (a) SLM, (b) LENS and (c–d) WAAM methods.

sample (Table 2). In contrast, SLM-fabricated titanium received much revealed that SLM-produced titanium has the highest yield strength (σy )
lower heat input (LED) and this stimulates significantly higher cooling and ultimate tensile strength (σUTS ) which can be attributed to its finest
rates, leading to the martensitic β to α′ phase transformation with a grain size and α′ phase. As mentioned before, the fine grain size and αʹ
reduced grain size, which has been also reported before [11,21]. phase obtained in SLM-produced CP-Ti can be related to a high cooling
The tensile stress-strain curves of the titanium samples are shown in rate created during the SLM process which promoted the β to αʹ phase
Fig. 2. All titanium samples possess excellent ductility, showing the transformation during fast cooling. LENS-produced titanium showed
maximum for the WAAM-produced titanium. Tensile properties (Fig. 3) slightly higher yield and ultimate strengths compared to those of
WAAM-produced titanium due to the smaller grain size and finer mi-
crostructural phase feature. Nonetheless, the α phase in CP-Ti obtained
using LENS and WAAM was also responsible for the resulting, lower
yield and ultimate strengths compared to those of SLM-produced CP-Ti
because the α phase usually has a lower strength and hardness than αʹ.
Additionally, it is well-known that the yield strength is inversely pro-
portional to the α grain size (or colony size) [22]. Generally, a smaller
yield strength can serve as a warning message for maintenance or in-
tervention before catastrophic failure happens in a potential implant
material [24] and improved yield strength can enhance an implant
material's resistance to permanent shape change, which may increase
its effectiveness for implant applications [25]. All AM-fabricated tita-
nium samples showed yield and ultimate strengths superior to those of
CP-Ti (Grades 1 and 2) and comparable to those of CP-Ti (Grade 3)
produced previously by casting [26,27]. Although yield and ultimate
strengths of AM-produced samples are relatively low compared to those
of powder metallurgy (P/M) produced CP-Ti [28], they showed sig-
nificantly superior ductility, indicating that AM-produced samples have
no major defects. Nonetheless, the ductility of the SLM and LENS-pro-
Fig. 2. Nominal tensile stress-strain curves of CP-Ti produced by SLM, LENS
and WAAM methods. duced samples narrowly fell below that of CP-Ti (Grade 1) but the

341
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

Fig. 4. SEM images of the fracture surfaces of CP-Ti after tensile testing: (a)
WAAM sample and (b) SLM sample.

SLM samples are consistent with their corresponding grain sizes which
is similar to previously reported findings [30,31]. As seen in Table 2,
Fig. 3. Comparison of the mechanical properties of CP-Ti produced by various
AM methods in our study as well as by conventional manufacturing methods the average grain width of the WAAM sample is 11.9 μm which is larger
reported earlier [27–29]. than that of the SLM sample (1.8 μm). Furthermore, a large number of
microvoids are present across the fracture surface of the SLM sample
and it has been reported that a smaller grain size causes an increased
WAAM-produced titanium fully exceeded the tensile ductility require-
density of nucleation sites to initiate microvoid coalescence [32]. As
ments.
such, since the SLM sample in our study possesses the smallest grain
Fig. 3 (b) displays the elastic modulus and Vickers hardness of the
size, nucleation and growth of numerous microvoids and their eventual
AM-produced titanium samples along with those reported for cast, and
coalescences resulted in the initiation and propagation of microscopic
samples produced by powder metallurgy (P/M) CP-Ti [28,29]. The
cracks (Fig. 4 (b)), leading to a reduction in the strain to failure (duc-
elastic modulus obtained for AM-produced samples are very compar-
tility) of the SLM-produced CP-Ti compared to WAAM-produced
able, however, their hardness shows the same trend to that observed for
counterpart.
σy and σUTS . The highest hardness (229 Hv) was observed for the SLM
It has been reported that biomedical implant materials with high
sample due to its smaller grain size and the presence of martensitic
strength and low elastic modulus are required to tolerate heavy loads
phase, while the WAAM sample presented the lowest value (190 Hv).
[29,33]. Biomedical materials are generally exposed to service condi-
However, it should be noted that all the AM-produced titanium samples
tions within the elastic region; therefore it is necessary to evaluate the
showed higher hardness than those of the cast and P/M CP-Ti.
elastic energy to ensure a safe performance during service. Fig. 5 (a)
As observed, the ductility of AM-produced CP-Ti samples was
shows a schematic representation of elastic energy in a typical tensile
comparable to conventionally produced CP-Ti. Nevertheless, SLM-pro-
stress-strain curve. The elastic energy can be described using the fol-
duced and WAAM-produced CP-Ti showed a difference of around 10%
lowing equation [33,34]:
ductility. Therefore, the fracture surfaces of these two samples were
studied by SEM and the results are presented in Fig. 4. The fracture 1 σy2
surface of the WAAM sample (Fig. 4(a)) shows mainly a mixture of δe = ⋅εe σy =
2 2E (3)
coarse and fine dimples, indicating a ductile-type fracture behavior.
Isolated and randomly distributed microvoids are also present across its Where δ is the elastic energy, εe is the elastic strain, σy and E are the
fracture surface. In contrast, the fracture surface of the SLM sample is yield strength and Young's modulus, respectively. Fig. 5 (b) presents the
mainly characterized by very fine dimples and microcracks. The elastic energy values of the CP-Ti samples produced by AM methods as
varying size of the dimples on the fracture surfaces of the WAAM and well as those of the cast and P/M CP-Ti calculated from the corre-
sponding mechanical properties reported in the literature [27,28]. As

342
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

Fig. 6. H/E and H3/E2 ratios and wear weight loss values for CP-Ti samples
produced by SLM, LENS and WAAM methods.

the yield pressure, can also be used to assess the wear resistance of a
material [11]. These aforementioned ratios have been used successfully
before to evaluate the wear performance of titanium materials [35,37],
so the values of H/E and H3/E2 ratios were measured for CP-Ti samples
in our study in order to evaluate the wear performance of the materials.
To this end, the Vickers microhardness values of the AM-produced ti-
tanium samples were converted to the unit of GPa using the equation
reported in the literature [38]. Fig. 6 also presents the H/E and H3/E2
ratios of the CP-Ti samples produced by the different AM methods. It is
evident that SLM-produced CP-Ti shows the maximum values for H/E
and H3/E2 which is in good agreement with its minimum weight loss.
The SEM images of titanium samples after wear testing are pre-
sented in Fig. 7. The wear mechanisms observed in all samples are si-
milar to those of other titanium materials in which wear occurs mainly
by abrasion [39,40]. The samples have almost identical composition,
and for this reason, the wear tracks of the AM-produced CP-Ti samples
show relatively similar features including ploughing grooves, micro-
cracks and delamination. However, the incidents of delamination ob-
served in SLM-produced titanium are less frequent than that observed
in the samples produced by LENS and WAAM as higher hardness and
Fig. 5. (a) A schematic illustration of elastic energy in a tensile stress-strain
strength was obtained from the samples produced by SLM. There was
curve, and (b) elastic energy versus elastic modulus for CP-Ti produced by
also evidence to suggest that initial cracking could not easily propagate
various AM methods in our study as well as those of CP-Ti reported previously
for samples produced by casting and conventional powder metallurgy (P/M)
into complete delamination of the surface, indicating higher wear re-
[27,28]. sistance in SLM-produced titanium. Generally, wear volume is inversely
proportional to the hardness of a given material according to Archard's
wear equation [41], therefore higher surface and subsurface plastic
can be seen, the elastic energy of the SLM-produced samples is higher
deformation occurs during wear for lower hardness and lower strength
than CP-Ti produced by LENS and WAAM. It is also greater than the
materials, causing deeper cracks and delamination. In this study, LENS
elastic energy of Grades 1, 2 and 3 CP-Ti. However, it is slightly lower
and WAAM-produced titanium showed higher weight loss as well as
than CP-Ti (Grade 4) and P/M CP-Ti due to the lower yield strength
deeper delamination and cracking consistent with these explanations.
observed in SLM-produced CP-Ti. These results show that for the pro-
Based on the results obtained in the present study, the samples
cess parameters and experimental conditions adopted in this study the
produced by SLM exhibited better wear performance and strength
SLM-produced CP-Ti possesses a better heavy-loading capability com-
compared to those produced using LENS and WAAM methods. This is
pared to CP-Ti produced by LENS and WAAM as well as the results
because the lower LED associated with SLM produces faster cooling
presented in the literature for CP-Ti produced by conventional casting,
rates and finer microstructures. However, it is important to note that
owing to the significant improvement in yield strength associated with
LED is varied with processing conditions; hence, cooling rates and mi-
samples produced by the SLM technique.
crostructures can also be modified in all AM processes which may
Weight loss values of the AM-produced titanium samples after wear
produce different results. For example, although the WAAM processing
testing are shown in Fig. 6. It is evident that titanium samples produced
parameters used in this study produced a LED of 140 J/mm, others have
by SLM show the lowest weight loss, suggesting that SLM provides
used different parameters in WAAM that produced much lower LEDs as
better wear resistance than other AM methods. Classical theories of
low as 147 J/mm [42]. Therefore while the present study has shown
wear show that hardness is a crucial factor in controlling the wear re-
that the improved properties of the SLM samples were associated with
sistance of a material, and a hard material usually presents better wear
the high cooling rates it is possible that modifications to the LENS and
resistance [35,36]. Therefore, the lowest weight loss observed in SLM-
WAAM such as the scale of feedstock and deposition process may result
produced CP-Ti can be associated with its higher hardness. In fact, it has
in higher cooling rates and thus finer α microstructures with better
been realized that the wear resistance of a given material is related to
strength (albeit with reduced ductility). All deposition pathways were
the elastic strain to failure which can be described by the hardness to
capable of producing high density preforms and should be considered
elastic modulus ratio (H/E) [37]. Another ratio (H3/E2), referred to as
for future research as potential manufacturing routes for biomedical

343
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

LED for WAAM, whereas the lowest LED was observed during SLM,
leading to the highest cooling rates and, consequently, the finest grain
size and a microstructure consisting of the martensitic (αʹ) phase.
Tensile testing of titanium produced by SLM showed the highest yield
and ultimate strengths, while WAAM-produced titanium demonstrated
the maximum ductility. Nonetheless, all the samples showed mechan-
ical properties at least comparable with those obtained previously from
conventional manufacturing techniques. There was a direct relationship
between wear performance and hardness of the titanium samples, and
the greatest wear resistance was observed in the SLM-produced tita-
nium. Furthermore, the ratios of H/E and H3/E2 measured for CP-Ti
produced by each AM method showed consistency with the corre-
sponding weight loss value, indicating that these ratios are good in-
dicators of wear resistance and should be further considered in future
investigations. Although CP-Ti produced by AM methods with the de-
position parameters selected for this study provided mechanical prop-
erties better than those of produced by conventional methods due to the
inherently higher cooling rates, it is suggested that SLM can provide
greater mechanical strength and wear properties compared to WAAM
and LENS which may be more desirable for biomedical applications.

Acknowledgments

The authors acknowledge the support of Queensland Centre for


Advanced Materials Processing and Manufacturing (AMPAM). Authors
also acknowledge the Australian Research Council (ARC) Hub of
transforming Australian industry through Additive Manufacturing
(IH130100008). M.J Bermingham acknowledges the support of the
Australian Research Council through DE160100260.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://


doi.org/10.1016/j.msea.2019.06.024.

References

[1] Q. Chen, G.A. Thouas, Metallic implant biomaterials, Mater. Sci. Eng. R Rep. 87
(2015) 1–57.
[2] I. Okulov, A. Volegov, H. Attar, M. Bönisch, S. Ehtemam-Haghighi, M. Calin,
J. Eckert, Composition optimization of low modulus and high-strength TiNb-based
alloys for biomedical applications, J. Mech. Behav. Biomed. Mater. 65 (2017)
866–871.
[3] I. Okulov, M. Bönisch, A. Okulov, A. Volegov, H. Attar, S. Ehtemam-Haghighi,
M. Calin, Z. Wang, A. Hohenwarter, I. Kaban, Phase formation, microstructure and
deformation behavior of heavily alloyed TiNb-and TiV-based titanium alloys,
Mater. Sci. Eng. A 733 (2018) 80–86.
[4] M. Geetha, A.K. Singh, R. Asokamani, A.K. Gogia, Ti based biomaterials, the ulti-
mate choice for orthopaedic implants–a review, Prog. Mater. Sci. 54 (2009)
397–425.
[5] M. Niinomi, M. Nakai, J. Hieda, Development of new metallic alloys for biomedical
applications, Acta Biomater. 8 (2012) 3888–3903.
[6] S. Ehtemam-Haghighi, Y. Liu, G. Cao, L.-C. Zhang, Influence of Nb on the β→ α
″martensitic phase transformation and properties of the newly designed Ti–Fe–Nb
Fig. 7. SEM micrographs showing the features of the worn surface of the tita- alloys, Mater. Sci. Eng. C 60 (2016) 503–510.
nium samples produced by (a) SLM, (b) LENS and (c) WAAM methods. [7] 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 bio-
medical applications, J. Alloy. Comp. 787 (2019) 570–577.
applications. Nevertheless, other important considerations ae required [8] Z. Zhang, S. Qu, A. Feng, X. Hu, J. Shen, Microstructural mechanisms during
multidirectional isothermal forging of as-cast Ti-6Al-4V alloy with an initial la-
given that powder-bed SLM is likely to offer greater precision and de- mellar microstructure, J. Alloy. Comp. 773 (2019) 277–287.
sign flexibility over LENS and WAAM and thus remains the most pro- [9] L. Bolzoni, E.M. Ruiz-Navas, E. Gordo, Quantifying the properties of low-cost
mising route to produce complex, near net shaped titanium components powder metallurgy titanium alloys, Mater. Sci. Eng. A 687 (2017) 47–53.
[10] L. Deng, S. Wang, P. Wang, U. Kühn, S. Pauly, Selective laser melting of a Ti-based
for biomedical application.
bulk metallic glass, Mater. Lett. 212 (2018) 346–349.
[11] H. Attar, S. Ehtemam-Haghighi, D. Kent, I. Okulov, H. Wendrock, M. Bӧnisch,
A. Volegov, M. Calin, J. Eckert, M. Dargusch, Nanoindentation and wear properties
4. Conclusions of Ti and Ti-TiB composite materials produced by selective laser melting, Mater. Sci.
Eng. A 688 (2017) 20–26.
[12] W. Chen, C. Chen, X. Zi, X. Cheng, X. Zhang, Y.C. Lin, K. Zhou, Controlling the
In this work, CP-Ti samples were fabricated using the different AM microstructure and mechanical properties of a metastable β titanium alloy by se-
methods of SLM, LENS and WAAM. A fine α′ microstructure was ob- lective laser melting, Mater. Sci. Eng. A 726 (2018) 240–250.
[13] A. Ataee, Y. Li, M. Brandt, C. Wen, Ultrahigh-strength titanium gyroid scaffolds
tained from SLM which was different to α microstructure obtained from
manufactured by selective laser melting (SLM) for bone implant applications, Acta
LENS and WAAM. Processing conditions provided significantly high

344
H. Attar, et al. Materials Science & Engineering A 760 (2019) 339–345

Mater. 158 (2018) 354–368. [28] L. Bolzoni, E.M. Ruiz-Navas, E. Gordo, Powder metallurgy CP-Ti performances:
[14] D. Zhao, Y. Huang, Y. Ao, C. Han, Q. Wang, Y. Li, J. Liu, Q. Wei, Z. Zhang, Effect of hydride–dehydride vs. sponge, Mater. Des. 60 (2014) 226–232.
pore geometry on the fatigue properties and cell affinity of porous titanium scaf- [29] P. Wang, Y. Feng, F. Liu, L. Wu, S. Guan, Microstructure and mechanical properties
folds fabricated by selective laser melting, J. Mech. Behav. Biomed. Mater. 88 of Ti–Zr–Cr biomedical alloys, Mater. Sci. Eng. C 51 (2015) 148–152.
(2018) 478–487. [30] M.-H. Cai, C.-Y. Lee, Y.-K. Lee, Effect of grain size on tensile properties of fine-
[15] J. de Krijger, C. Rans, B. Van Hooreweder, K. Lietaert, B. Pouran, A.A. Zadpoor, grained metastable β titanium alloys fabricated by stress-induced martensite and its
Effects of applied stress ratio on the fatigue behavior of additively manufactured reverse transformations, Scripta Mater. 66 (2012) 606–609.
porous biomaterials under compressive loading, J. Mech. Behav. Biomed. Mater. 70 [31] Y.G. Ko, D.H. Shin, K.-T. Park, C.S. Lee, An analysis of the strain hardening behavior
(2017) 7–16. of ultra-fine grain pure titanium, Scripta Mater. 54 (2006) 1785–1789.
[16] W.E. Frazier, Metal additive manufacturing: a review, J. Mater. Eng. Perform. 23 [32] M. Bermingham, L. Nicastro, D. Kent, Y. Chen, M. Dargusch, Optimising the me-
(2014) 1917–1928. chanical properties of Ti-6Al-4V components produced by wire+ arc additive
[17] H. Gong, K. Rafi, H. Gu, G.J. Ram, T. Starr, B. Stucker, Influence of defects on manufacturing with post-process heat treatments, J. Alloy. Comp. 753 (2018)
mechanical properties of Ti–6Al–4 V components produced by selective laser 247–255.
melting and electron beam melting, Mater. Des. 86 (2015) 545–554. [33] S. Ehtemam-Haghighi, Y. Liu, G. Cao, L.-C. Zhang, Phase transition, microstructural
[18] L. Murr, S. Quinones, S. Gaytan, M. Lopez, A. Rodela, E. Martinez, D. Hernandez, evolution and mechanical properties of Ti-Nb-Fe alloys induced by Fe addition,
E. Martinez, F. Medina, R. Wicker, Microstructure and mechanical behavior of Mater. Des. 97 (2016) 279–286.
Ti–6Al–4V produced by rapid-layer manufacturing, for biomedical applications, J. [34] M. Clifford, K. Simmons, P. Shipway, An Introduction to Mechanical Engineering,
Mech. Behav. Biomed. Mater. 2 (2009) 20–32. CRC Press, 2012.
[19] M. Bermingham, S. McDonald, M. Dargusch, Effect of trace lanthanum hexaboride [35] S. Ehtemam-Haghighi, G. Cao, L.-C. Zhang, Nanoindentation study of mechanical
and boron additions on microstructure, tensile properties and anisotropy of Ti-6Al- properties of Ti based alloys with Fe and Ta additions, J. Alloy. Comp. 692 (2017)
4V produced by additive manufacturing, Mater. Sci. Eng. A 719 (2018) 1–11. 892–897.
[20] H. Attar, S. Ehtemam-Haghighi, D. Kent, X. Wu, M.S. Dargusch, Comparative study [36] J. Xu, G. dong Wang, X. Lu, L. Liu, P. Munroe, Z.-H. Xie, Mechanical and corrosion-
of commercially pure titanium produced by laser engineered net shaping, selective resistant properties of Ti–Nb–Si–N nanocomposite films prepared by a double glow
laser melting and casting processes, Mater. Sci. Eng. A 705 (2017) 385–393. discharge plasma technique, Ceram. Int. 40 (2014) 8621–8630.
[21] D. Gu, Y.-C. Hagedorn, W. Meiners, G. Meng, R.J.S. Batista, K. Wissenbach, [37] A. Hynowska, E. Pellicer, J. Fornell, S. González, N. van Steenberge, S. Suriñach,
R. Poprawe, Densification behavior, microstructure evolution, and wear perfor- A. Gebert, M. Calin, J. Eckert, M.D. Baró, Nanostructured β-phase Ti–31.0 Fe–9.0 Sn
mance of selective laser melting processed commercially pure titanium, Acta Mater. and sub-μm structured Ti–39.3 Nb–13.3 Zr–10.7 Ta alloys for biomedical applica-
60 (2012) 3849–3860. tions: microstructure benefits on the mechanical and corrosion performances,
[22] M.J. Donachie, Titanium: a Technical Guide, ASM international, 2000. Mater. Sci. Eng. C 32 (2012) 2418–2425.
[23] S. Bontha, N.W. Klingbeil, P.A. Kobryn, H.L. Fraser, Thermal process maps for [38] Y. Sun, J. Liang, Z.-H. Xu, G. Wang, X. Li, Nanoindentation for measuring individual
predicting solidification microstructure in laser fabrication of thin-wall structures, phase mechanical properties of lead free solder alloy, J. Mater. Sci. Mater. Electron.
J. Mater. Process. Technol. 178 (2006) 135–142. 19 (2008) 514–521.
[24] P.F. Santos, M. Niinomi, H. Liu, K. Cho, M. Nakai, A. Trenggono, S. Champagne, [39] S. Ehtemam-Haghighi, K. Prashanth, H. Attar, A.K. Chaubey, G. Cao, L. Zhang,
H. Hermawan, T. Narushima, Improvement of microstructure, mechanical and Evaluation of mechanical and wear properties of Ti xNb 7Fe alloys designed for
corrosion properties of biomedical Ti-Mn alloys by Mo addition, Mater. Des. 110 biomedical applications, Mater. Des. 111 (2016) 592–599.
(2016) 414–424. [40] G. Straffelini, A. Molinari, Dry sliding wear of Ti–6Al–4V alloy as influenced by the
[25] H. Attar, S. Ehtemam-Haghighi, D. Kent, M.S. Dargusch, Recent developments and counterface and sliding conditions, Wear 236 (1999) 328–338.
opportunities in additive manufacturing of titanium-based matrix composites: a [41] Q. Wang, P.-Z. Zhang, D.-B. Wei, X.-H. Chen, R.-N. Wang, H.-Y. Wang, K.-T. Feng,
review, Int. J. Mach. Tool Manuf. 133 (2018) 85–102. Microstructure and sliding wear behavior of pure titanium surface modified by
[26] J.D. Beal, R. Boyer, D. Sanders, Forming of titanium and titanium alloys, double-glow plasma surface alloying with Nb, Mater. Des. 52 (2013) 265–273.
Metalwork.: Sheet Form.(ASM Handbook) 14 (2006) 656–669. [42] Q. Wu, Z. Ma, G. Chen, C. Liu, D. Ma, S. Ma, Obtaining fine microstructure and
[27] M. Niinomi, Mechanical properties of biomedical titanium alloys, Mater. Sci. Eng. A unsupported overhangs by low heat input pulse arc additive manufacturing, J.
243 (1998) 231–236. Manuf. Process. 27 (2017) 198–206.

345

You might also like