Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science & Engineering A 790 (2020) 139695

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Compressive fatigue properties of additive-manufactured Ti-6Al-4V cellular


material with different porosities
Ming-Wei Wu *, Jhewn-Kuang Chen , Bo-Huan Lin , Po-Hsing Chiang , Mo-Kai Tsai
Department of Materials and Mineral Resources Engineering, National Taipei University of Technology, No. 1, Sec. 3, Zhong-Xiao E. Rd., Taipei, 10608, Taiwan

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

Keywords: Selective laser melting (SLM) is a novel additive manufacturing (AM) technique for producing cellular metallic
Selective laser melting materials with designed pore structures. SLM Ti-6Al-4V cellular alloys have been extensively studied for
Ti-6Al-4V biomedical and other applications. The objective of this study was to investigate the effects of porosity (33 vol%,
Cellular material
50 vol%, and 84 vol%) on the compressive fatigue performance and fracture mechanism of an SLM Ti-6Al-4V
Hot isostatic pressing
Fatigue
cellular material with a new cuboctahedron COH-Z unit cell. Furthermore, the influences of hot isostatic
Fracture pressing (HIP) at 1000 � C/150 MPa on the fatigue properties and fracture were also examined in this study.
The results showed that the fatigue strain slowly accumulated when the stress cycle was below the 70% fatigue
life. After the 70% fatigue life, the ratcheting strain rate (dε/dN) gradually increased. A shear band along a
direction inclined at ~45� to the stress axis could be definitely identified at about 95% fatigue life, as demon­
strated by the digital image correlation technique. With further increases in the stress cycle, the shear band was
intensified and finally led to rapid fatigue failure. The HIP process did not alter the fracture behaviors of the SLM
cellular alloy, although the fatigue life was much improved by several times. On the other hand, the fatigue
endurance ratio of the SLM cellular alloy with 33 vol% was 0.5. Raising the porosity from 33 vol% to 84 vol%
reduced the endurance ratio from 0.5 to 0.15 due mainly to the high notch sensitivity of α0 -martensite in the SLM
cellular alloy. The HIP treatment changed the microstructure from α0 -martensite to lamellar αþβ phases and thus
significantly improved the fatigue endurance performance. With the combination of the COH-Z unit cell and HIP
treatment, the endurance ratios at 106 cycles of the Ti-6Al-4V cellular alloy with porosities of 33 vol% to 84 vol%
were as high as 0.5. This superior endurance ratio is advantageous to the long-term application of SLM Ti-6Al-4V
cellular alloy.

1. Introduction The unit cell of the cellular structure dominates the mechanical
properties of an AM cellular solid [14,15,18–20]. Zhao et al. [20]
Ti-based cellular alloys, particularly Ti-6Al-4V, have been exten­ examined the fatigue strength of an EBM Ti-6Al-4V cellular solid with
sively produced and studied for application as biomedical materials cubic, rhombic dodecahedron, and G7 structures and reported that the
[1–4]. Recently, additive manufacturing (AM) has been acknowledged ratcheting strain rate (dε/dN) is the key factor affecting fatigue perfor­
as a versatile technique to produce Ti-based alloys with controlled mance. Increasing the buckling deformation on the strut can lower the
porosity and shapes layer-by-layer [5–12]. AM Ti-based alloys have high cyclic ratcheting rate and improve the fatigue properties. The fatigue
potential for future application as high-performance materials in the performances of the various cell structures increase in the order of G7,
biomedical and aerospace fields. Among the various AM processes, se­ rhombic dodecahedron, and cubic unit cells. However, Li et al. [15]
lective laser melting (SLM) and electron beam melting (EBM) are the indicated that increasing the buckling deformation on the strut leads to
most frequently applied methods of fabricating Ti-based cellular solids brittle fracture under uniaxial compressive stress. Thus, cellular AM
[13–20]. For biomaterials, the fatigue properties of the materials are Ti-6Al-4V with cubic and rhombic dodecahedron structures exhibited
extremely important due to their loading states and long-term use. The brittle characteristics and a high Young’s modulus under compressive
compressive fatigue performance of AM Ti-6Al-4V cellular materials has stress [15]. These findings clearly show that it is very difficult to achieve
been examined in numerous studies [19–24]. an AM porous Ti-6Al-4V material with the excellent combination of a

* Corresponding author.
E-mail address: mwwu@ntut.edu.tw (M.-W. Wu).

https://doi.org/10.1016/j.msea.2020.139695
Received 31 January 2020; Received in revised form 28 May 2020; Accepted 31 May 2020
Available online 6 June 2020
0921-5093/© 2020 Elsevier B.V. All rights reserved.
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

low Young’s modulus, a high fatigue resistance, and sufficient ductility. the new cuboctahedron structure, cylindrical specimens with a sand­
The mechanical properties of SLM Ti-6Al-4V alloys are generally wiched architecture were prepared, as illustrated in Fig. 2(a). The
inferior due to the low-ductility microstructure of α0 -martensite [12, diameter and height of the cylinder were 13 mm and 25 mm, respec­
23–29]. Therefore, these alloys are heat treated to improve the me­ tively. The heights of the solid top section, cellular middle section, and
chanical performances by modifying the microstructure [23,24,27,29]. solid bottom section were 5 mm, 15 mm, and 5 mm, respectively. The
Hooreweder et al. [23] investigated the effects of stress relieving and hot strut width was adjusted to build Ti-6Al-4V cellular materials with three
isostatic pressing (HIP) on the compressive fatigue properties of an SLM different porosities (33 vol%, 50 vol%, and 84 vol%) between the solid
Ti-6Al-4V cellular alloy. They found that both stress relieving and HIP top and bottom sections.
treatment can effectively improve the fatigue properties under a high A spherical Ti-6Al-4V powder with a median particle size of 34 μm
cycle fatigue regime. Wu et al. [24] reported that HIP treatment at 1000 and a selective laser melting machine (SLM 250 HL, SLM Solutions

C/150 MPa significantly increases the fatigue properties of SLM GmbH, Lübeck, Germany) were used to produce the sandwiched speci­
Ti-6Al-4V cellular material. mens investigated in this study. The samples were prepared under a
Other than the microstructure and the type of unit cell, the porosity protective atmosphere of argon during the SLM process. The details of
of the AM metallic material also obviously affects the fatigue properties. the experimental parameters for the preparation of the specimens can be
The fatigue life can be greatly reduced by increases in porosity [8,19,
22]. The porosity of the cellular material can be varied depending on the
application. Thus, the main purpose of this study was to identify the
influences of porosity (33 vol%, 50 vol%, and 84 vol%) on the
compressive fatigue properties and fracture behavior of an SLM
Ti-6Al-4V cellular solid with a specific cuboctahedron unit cell, which
was designated as a COH-Z structure in a previous study [29]. Digital
image correlation (DIC) has recently been used to analyze the strain
distribution on the struts of AM cellular metallic materials [30–34]. In
this study, DIC was applied to examine the progress of the fatigue strain
distribution as a function of stress cycle. Furthermore, previous studies
[23,24,29] have indicated that the fatigue performance of SLM
Ti-6Al-4V cellular alloy can be improved by adequate heat treatment.
The second objective was to investigate the role of HIP on fatigue per­
formance. In general, the fatigue endurance ratios at 106 cycles of AM
Ti-based cellular alloys are below 0.2 [8,21,22]. This low endurance
ratio could prohibit the application of the AM Ti-6Al-4V cellular alloys.
Thus, the relationships between porosity, microstructure, and fatigue
endurance ratio are discussed in this study.

2. Experimental procedure

Fig. 1 presents the schematics of the cuboctahedron unit cell (COH-Z)


proposed in a previous study [29]. This unit cell is based on the tradi­
tional cuboctahedron structure. Three additional straight struts along
the X, Y, and Z axes and four diagonal struts were added into the
cuboctahedron structure. The dimensions of the COH-Z structure in the
X, Y, and Z axes were 2 mm, 2 mm, and 1.44 mm, respectively. To
investigate the effect of porosity on the compressive fatigue properties of

Fig. 2. The schematics of the sandwiched cylindrical specimens used for the
compressive fatigue test. (a) original appearance (b) after preparation for in-situ
Fig. 1. The unit cell of the new cuboctahedron COH-Z structure. observation.

2
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

found elsewhere [29]. HIP has been found to obviously improve the
brittleness of the SLM Ti-6Al-4V cellular solid in a previous study [24].
Therefore, some SLM specimens were HIPed, followed by furnace
cooling in a hot isostatic press made by Avure Technologies. The pres­
sure, temperature, and holding time were respectively 150 MPa, 1000

C, and 60 min in the HIP process. Argon was the pressurized gas used in
the HIP treatment.
For the microstructural observations, the SLM and HIP specimens
were ground, polished, and then chemically etched. The etchant was
composed of 5 mL HF, 25 mL HNO3, and 50 mL distilled water. The
metallographic samples were finally analyzed under an optical micro­
scope and a scanning electron microscope (SEM, JSM-6510, JEOL,
Tokyo, Japan). In addition, X-ray computed tomography (CT, MET­
ROTOM 800, Carl Zeiss Industrielle Messtechnik GmbH, Oberkochen,
Germany) was used to estimate the strut porosity. The voxel size used
was 21 μm. The other operating parameters of the X-ray CT experiments
have been reported in a previous study [24].
To identify the compressive fatigue properties of the SLM and HIP
specimens with different porosities, a fatigue testing machine (EHF-
EM100kN-10L) with a Shimadzu 4830 Servo Controller was utilized as
per ASTM standard E466. The stress ratio (R) and working frequency
were 0.1 and 10 Hz, respectively. A sinusoidal loading wave was used
during the fatigue test. The criterion for fatigue fracture was 5% strain.
The Stress-Number (S-N) curve was analyzed to obtain the fatigue
strength at 106 cycles. The fatigue tests were executed in the ranges of
15%–80% of the yield strength for each material. The fatigue fracture
surfaces of the SLM and HIP specimens were observed under SEM.
Before the fatigue test, a universal material testing machine (HT-9510,
Hung Ta Instrument Co., Taiwan) was also used to analyze the yield
strength under uniaxial compressive force. The loading rate was 0.125
mm/min.
To capture the progress of the fatigue failure, the original cylindrical
specimens were ground and polished. The polished samples were then
observed using a CCD camera during the compressive fatigue test for in-
situ observation. The schematics of the polished specimen for the fatigue
test are shown in Fig. 2(b). Furthermore, the successive images taken
from the in-situ observation were analyzed with two-dimensional DIC
(Vic-2D, Correlated Solutions, Inc., USA) software to obtain the strain
distributions on the observed plane as a function of fatigue cycles. The Fig. 3. The microstructures of the Ti-6Al-4V cellular alloy. (a) SLM (b) HIP.
evolutions of the deformation and fracture during the fatigue test were
clarified according to the results from in-situ observation and DIC.
Representative results are presented in this study. Table 1
The yield strengths of the SLM and HIP specimens as a function of porosity.

3. Results and discussion Condition Porosity (vol %) Yield Strength (MPa)

SLM 33 222
3.1. Microstructure and uniaxial compressive property SLM 50 143
SLM 84 47
HIP 33 195
The microstructures of the SLM and HIP samples are shown in Fig. 3. HIP 50 100
The results indicated that the HIP process sufficiently modified the HIP 84 25
microstructure of the SLM sample. The SLM specimen consisted of
acicular α0 -martensite dispersed within columnar grains, as shown in
Fig. 3(a). After HIP at 1000 � C/150 MPa, the microstructure was yield strength due to the microstructural change from α0 -martensite to
completely transformed to fully lamellar αþβ phases in equiaxed grains. αþβ dual phases. After HIP, the microhardness of the struts was lower,
This microstructural change has been reported in several studies [23,24, falling from 403 HV to 324 HV [24].
27,29]. The strut porosity of the medium cellular layer was analyzed
using X-ray CT. The strut porosities of the SLM samples with 33 vol%, 50 3.2. Stress-number curve
vol%, and 84 vol% porosity were 0.0617 vol%, 0.0101 vol%, and
0.0027 vol%, respectively. After HIP, the strut porosities were decreased The compressive fatigue S-N curves of the SLM and HIP specimens
to less than 0.001 vol%. Therefore, the residual pores in the struts were with three porosities are presented in Fig. 4. The allowable cyclic stress
definitely reduced due to the densification effect of HIP treatment. at a given fatigue cycle obviously decreased with increases in the
The yield strengths of the SLM and HIP specimens with three po­ porosity. Fig. 5 shows the fatigue strengths at 106 cycles of the SLM and
rosities are listed in Table 1. Increasing the porosity of the SLM speci­ HIP samples. The fatigue strengths of the SLM specimens with 33 vol%,
mens from 33 vol% to 84 vol% apparently decreased the yield strength 50 vol%, and 84 vol% porosity were 111 MPa, 43 MPa, and 7 MPa,
from 222 MPa to 47 MPa. Moreover, the yield strengths of the HIP respectively. Moreover, the fatigue strengths of the HIP samples with 33
samples decreased from 195 MPa to 25 MPa when the porosity was vol%, 50 vol%, and 84 vol% porosity were 98 MPa, 55 MPa, and 11 MPa,
increased from 33 vol% to 84 vol%. The HIP treatment decreased the respectively. As shown in section 3.1, the yield strength of an SLM

3
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

Fig. 4. The S-N curves of SLM and HIP samples with porosities of 33 vol%, 50
vol%, and 84 vol%.

Fig. 6. The fatigue fracture surfaces of an SLM sample with 50% porosity.

Fig. 5. The fatigue strengths at 106 cycles of SLM and HIP samples as a function Figs. 6(b) and 7(b). Straight struts along the X, Y, and Z axes were
of porosity. broken, as indicated by the dashed arrows. Diagonal struts were also
fractured at the intersection between the diagonal struts and the central
sample was always higher than that of an HIP specimen at the same nodes, as indicated by the solid arrows. Fatigue striation was evident on
porosity. When the porosity was 33 vol%, the fatigue strength at 106 the fracture surfaces, as shown in Figs. 6(c) and 7(c).
cycles of the SLM sample was higher that of the HIP specimen. However, Moreover, the polished surfaces of the SLM and HIP specimens after
when the porosity was 50 vol% or 84 vol%, the HIP process increased fatigue failure were also examined by SEM to identify the fracture paths,
the fatigue strengths at 106 cycles of the SLM specimens, even though and the SEM images are shown in Fig. 8. Cracking at ~45� to the loading
the yield strength of the HIP specimen was lower. axis was apparent in both the SLM and HIP specimens. Fig. 8 also shows
that several straight struts along the X, Y, and Z axes, indicated by the
3.3. Fracture surface dashed arrows, were broken after fatigue test. Some diagonal struts,
indicated by the solid arrows, were ruptured at the junctions of the di­
Figs. 6 and 7 show the fatigue fracture surfaces of the SLM and HIP agonal struts and the central nodes. The fracture observations in Fig. 8
specimens, respectively. The results indicated that the fracture sites in completely matched the findings obtained from the fracture surfaces in
the SLM and HIP samples were identical, as can be seen from comparing Figs. 6 and 7. These previous fracture analyses implied that the HIP

4
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

Fig. 8. The fatigue fracture path of the Ti-6Al-4V cellular alloy with 50%
porosity after compressive fatigue fracture. (a) SLM (b) HIP.

be clearly discriminated from the pictures taken by the CCD camera


because the depth of field of the CCD camera was low.
From the previous in-situ observation, it was impossible to clarify the
exact fatigue fracture behavior. The strain mapping as a function of fa­
tigue cycle obtained by DIC is also presented in Fig. 9. The color bar on
the right side of Fig. 9 represents the relative levels of strain. The strain
mapping clearly indicated that obvious strain localization along a di­
rection inclined at ~45� to the fatigue stress axis occurred at 30,450
cycles. The shear band was extended and became clearer with increases
in the fatigue cycles, and the SLM specimen with 33 vol% porosity
finally fractured at 31,600 cycles. Therefore, shear fracture was the main
mechanism of fatigue fracture in the SLM sample with 33 vol% porosity.
Fig. 10 shows the photos and strain mappings at various fatigue
cycles of the SLM sample with 50 vol% porosity. Based on the photos
Fig. 7. The fatigue fracture surfaces of an HIP sample with 50% porosity.
taken at different numbers of fatigue cycles, no apparent strut defor­
mation or failure occurred when the specimen reached the criterion of
the fatigue fracture, 5% strain, at 9,889 cycles. However, the strain
treatment did not apparently change the dominant fracture site in the
mappings showed that strain localization at a direction inclined at ~45�
COH-Z unit cell.”
to the stress axis occurred at 9,700 cycles, and the sample eventually
fractured at 9,889 cycles. The results for the SLM sample with 84 vol%
3.4. Fatigue fracture evolution of the SLM specimen
porosity were similar to those of the SLM sample with 33 vol% and 50
vol% porosity and are not shown here.
Figs. 6–8 show the basic fatigue fracture behaviors of the SLM and
HIP specimens. To further clarify the evolution of the fatigue strain and
fracture, in-situ observation in combination with DIC was performed. 3.5. Fatigue fracture evolution of the HIP sample
The images taken at various fatigue cycles and the corresponding DIC
strain mappings of the SLM sample with 33 vol% porosity are presented Fig. 11 shows the progress of fatigue failure and strain distribution of
in Fig. 9. The images obtained from the in-situ observation indicated that an HIP sample with 50 vol% porosity. The images taken at different
no obvious strut deformation or fracture occurred when the fatigue cycle points in the stress cycles showed no obvious changes between 0 and
was below 26,400. Some areas, indicated by the arrows, were deformed 83,400 cycles. However, strut fracture or distortion, indicated by the
or fractured at 30,450 cycles. The SLM specimen with 33 vol% porosity arrows, occurred in a few areas when the number of fatigue cycles was
rapidly fractured at 31,600 cycles. Strut distortion or breakage could not increased to 92,700. The strain mappings demonstrated that the strain

5
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

Fig. 9. The in-situ images and corresponding DIC strain mappings of an SLM sample with 33 vol% porosity under cyclic compressive stress.

Fig. 10. The in-situ images and corresponding DIC strain mappings of an SLM sample with 50 vol% porosity under cyclic compressive stress.

Fig. 11. The in-situ images and corresponding DIC strain mappings of an HIP sample with 50 vol% porosity under cyclic compressive stress.

6
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

localization along a direction inclined at ~45� to the stress axis gradu­ designs. They showed that the main fracture mode of an SLM Ti-6Al-4V
ally evolved in the specimen. The shear band intensified and was cellular solid was shear fracture at about 45� to the stress axis [21–24,
extended with increases in the number of fatigue cycles, resulting in 35]. Zargarian et al. [35] indicated that the compressive fatigue failure
fatigue failure at 93,150 cycles. The findings for HIP samples with 33 vol was inclined at 45� to the stress direction in an SLM Ti-6Al-4V cellular
% or 84 vol% porosity were identical to those for the HIP sample with solid with diamond, rhombic dodecahedron, and truncated cuboctahe­
50 vol% porosity. dron unit cells by simulations of the finite element method. In this study,
Fig. 12 shows the influences of the stress cycles on the fatigue strains the compressive fatigue fracture modes of the SLM and HIP specimens
of the SLM and HIP specimens with the three porosities under a with the COH-Z unit cell were also shear fracture inclined at 45� to the
maximum fatigue stress of 70% yield strength. In general, the ratcheting stress direction, although the fatigue lives were obviously different.
strain rate was fairly stable when the stress cycles were lower than the Wu et al. [34] also investigated the uniaxial compressive properties
70% fatigue life, irrespective of the sample condition. However, when and fracture behaviors of SLM and HIP samples with the COH-Z unit cell.
the number of fatigue cycles exceeded the 70% fatigue life, the slope of They found that the struts were seriously deformed or broke at 50%
the curve deviated and gradually rose as the number of stress cycles compressive strain and that both the SLM and HIP samples exhibited a
increased. The increase in strain with the number of stress cycles finally layer-by-layer fracture mode under uniaxial compressive stress. How­
induced the shear band. According to the DIC strain mappings in ever, the compressive fatigue fracture modes of both the SLM and HIP
Figs. 9–11, an obvious shear band first appeared at ~95% fatigue life. samples were obviously different from the uniaxial compressive fracture
The generation of shear band further enhanced the ratcheting strain rate behavior. This discrepancy in the fracture mechanism could be ascribed
and led to rapid fatigue failure. to the stress mode. During uniaxial compressive test, the compressive
From the in-situ images in Figs. 9–11, the fatigue fracture behavior of strain was gradually increased from zero to 50%. In contrast, the strain
the SLM Ti-6Al-4V cellular material could not be definitely identified. of the struts was significantly lower under compressive fatigue stress
The struts were not seriously distorted or broken after the fatigue test than under uniaxial compressive stress. The struts were not seriously
because the criterion for fatigue fracture was 5% compressive strain. distorted or fractured, but only evolved into a shear band and fracture.
Some studies [30,31,34] have utilized the DIC method to identify the Thus, the fracture mechanism of SLM Ti-6Al-4V cellular material is
strain in AM cellular alloys and indicated that this technique is not directly related to the stress mode.
suitable for quantifying the strain in cellular alloys. However, the strain
distributions obtained by DIC in Figs. 9–11 clearly revealed the strain 3.6. Fatigue endurance performance
localization in the cellular alloy and helped to clarify the evolution of the
fatigue strain and fracture. Thus, DIC is still an effective analytical To further evaluate the effects of porosity and microstructure on
method for qualifying the strain distributions in cellular materials. fatigue endurance performance, the fatigue strength/yield strength ra­
Previous analyses of the fatigue deformation and fracture showed tios as a function of fatigue cycles of the SLM and HIP specimens are
that the SLM and HIP specimens with three porosities all exhibited shown in Fig. 13. The trend in Fig. 13 demonstrates that the fatigue
similar fracture modes. During fatigue testing, the shear strain along a strength/yield strength ratio decreased with increases in the porosity of
direction inclined at ~45� to the stress axis obviously dominated the the SLM specimen. Fig. 14 shows the fatigue endurance ratios at 106
fatigue strain and fracture behavior of the SLM and HIP specimens. cycles of the SLM and HIP samples with three porosities. The results
Cracking at ~45� to the loading axis was evident in the SEM observation indicate that the endurance ratios at 106 cycles of the SLM samples with
of the polished surface after fatigue fracture in Fig. 8. The diagonal shear 33 vol%, 50 vol%, and 84 vol% porosity were 0.5, 0.3, and 0.15,
band revealed by the DIC method in Figs. 9–11 well corresponded to the respectively. Moreover, the endurance ratios at 106 cycles of the HIP
findings in Fig. 8. It should be noted that at a given porosity, the fatigue samples with 33 vol%, 50 vol%, and 84 vol% porosity were respectively
life of the HIP sample was several times longer than that of the SLM
sample, even though the fracture modes of the two types were identical,
as shown in Figs. 10–12.
Numerous studies have investigated the compressive fatigue fracture
behavior of AM Ti-6Al-4V cellular materials with unit cells of various

Fig. 12. The fatigue strains of SLM and HIP specimens with three porosities as Fig. 13. The fatigue endurance ratios at 106 cycles of SLM and HIP samples
a function of fatigue cycle under a maximum cyclic stress of 70% yield strength. with porosities of 33 vol%, 50 vol%, and 84 vol%.

7
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

porosity of the HIP specimen from 33 vol% to 84 vol% degraded the


yield strength and fatigue strength but did not greatly reduce the fatigue
endurance performance, as shown in Table 1 and Figs. 5 and 14. Fig. 14
shows that the fatigue endurance ratio of the HIP specimen with 84 vol%
was still as high as 0.45. Thus, the fatigue endurance performances of
the cellular solids investigated in this study were dominated by the
microstructure. The porosity of the SLM Ti-6Al-4V cellular material was
obviously an important factor when the microstructure was brittle
α0 -martensite.
Yarari et al. [22] attributed the low fatigue endurance ratio of SLM
Ti-6Al-4V cellular alloy to four main factors: high surface roughness, the
high notch sensitivity of Ti-6Al-4V, low-thickness struts, and residual
thermal stress. When a selected area is scanned by a high-energy laser or
electron beam, heat is conducted to the neighboring regions, and many
raw particles can thus be sintered to the surface of the original selected
area, as shown in Fig. 6. Thus, high surface roughness is a normal feature
of AM cellular materials. Hooreweder et al. [23] showed that chemical
etching of SLM Ti-6Al-4V cellular alloy can produce a smooth surface
and reduced surface irregularities, resulting in improved fatigue prop­
erties at 106 cycles.
In this study, the raw powders sintered to the surfaces of the struts
were not eliminated by any post-treatment. The fatigue endurance ratio
of the SLM sample with 33 vol% porosity was 0.5, which is similar to the
fatigue endurance ratio at 106 cycles of fully-dense Ti-based alloys and
Ti-6Al-4V [36–40]. This previous finding demonstrates that the fatigue
Fig. 14. The fatigue endurance ratios at 106 cycles of SLM and HIP specimens
as a function of porosity. endurance performance of the SLM Ti-6Al-4V cellular material with the
new cuboctahedron COH-Z unit cell is high, although the endurance
ratio gradually decreased when the porosity was increased from 33 vol%
0.5, 0.55, and 0.45.
to 84 vol% due to the low ductility of α0 -martensite. The HIP treatment
The fatigue endurance ratios of AM Ti-6Al-4V cellular alloys are
modified the microstructure and improved the ductility of the SLM
generally low, seriously restricting their application. Yarari et al. [22]
sample, thereby alleviating the stress/strain concentration effects
reported that the fatigue endurance ratio at 106 cycles of an SLM
caused by the pores. Consequently, the fatigue endurance ratios of the
Ti-6Al-4V cellular alloy with a rhombic dodecahedron unit cell was
HIP specimens with porosities of 33 vol% to 84 vol% could be main­
about 0.12. The endurance ratio did not obviously change when the
tained at ~0.5. This superior fatigue endurance ratio over a wide range
porosity in the cellular alloy was modified from 68 vol% to 84 vol%. The
of porosity should be very beneficial to the design of SLM Ti-6Al-4V
fatigue endurance ratios at 2 � 105 cycles of an SLM Ti-6Al-4V cellular
cellular alloys for long-term application.
solid with truncated cuboctahedron, diamond, and rhombic dodecahe­
dron unit cells were ~0.35, 0.31, and 0.2, respectively [19]. Moreover,
4. Conclusions
the fatigue endurance ratios at 106 cycles of EBM Ti-6Al-4V and β-Ti
(Ti-23.9 wt%Nb-3.9 wt%Zr-8.2 wt%Sn-0.19 wt%O) cellular alloys with
1. The microstructure of the SLM samples was α0 -martensite. After HIP,
a rhombic dodecahedron unit cell and porosity of 75 vol% were ~0.1
the microstructure was transformed to lamellar αþβ phases.
and ~0.2, respectively [8,21].
Furthermore, the HIP treatment decreased the strut porosity of the
The findings in Fig. 14 demonstrate that the fatigue endurance ratio
SLM specimens to less than 0.001 vol%, as determined by X-ray CT.
of the SLM Ti-6Al-4V cellular solid with the COH-Z unit cell and 33%
2. During fatigue tests, the ratcheting strain rate as a function of fatigue
porosity was 0.5. Increasing the porosity of the SLM specimen from 33
cycle was slow when the stress cycle was lower than the 70% fatigue
vol% to 84 vol% significantly reduced the fatigue endurance ratio from
life. When the number of fatigue cycles exceeded the 70% fatigue
0.5 to 0.15. In contrast, the fatigue endurance ratios of the HIP samples
life, the strain rapidly increased with the number of stress cycles.
with the three porosities were near 0.5. The fatigue endurance ratios of
Strain localization along a direction inclined at ~45� to the stress
the HIP specimen were not reduced by increasing the porosity from 33
axis was observed at about 95% fatigue life. The shear band was
vol% to 84 vol%. This obvious difference in the fatigue endurance
extended with increases in the number of fatigue cycles and finally
property can be mainly attributed to the microstructure. The micro­
led to the high ratcheting strain rate and rapid fatigue failure. The
structures of the SLM and HIP specimens were α0 -martensite and
fracture modes of the SLM and HIP specimens were identical.
lamellar αþβ, respectively. Many studies have found that solid and
3. The fatigue endurance ratio of the SLM Ti-6Al-4V cellular material
cellular SLM Ti-6Al-4V alloy with α0 -martensite is relatively brittle [12,
with COH-Z unit cells and porosity of 33 vol% was 0.5. However,
23–29]. Increasing the porosity of the SLM Ti-6Al-4V cellular material
increasing the porosity from 33 vol% to 84 vol% reduced the
introduced more stress/strain concentrators. The fatigue stress/strain
endurance ratio from 0.5 to 0.15. This phenomenon can be attributed
could not easily be relieved by plastic deformation due to the low
to the low ductility of α0 -martensite, which resulted in high notch
ductility of α0 -martensite, which explains the deterioration of the fatigue
sensitivity.
endurance performance with increases in porosity.
4. The HIP process at 1000 � C/150 MPa transformed the microstructure
However, after HIP, α0 -martensite in the SLM specimen was trans­
to lamellar αþβ phases. This microstructural change significantly
formed to fully lamellar αþβ dual phases, which were similar to the
alleviated the stress/strain concentration effects and improved the
microstructure in the traditional cast and wrought Ti-6Al-4V alloy.
compressive fatigue performances. The fatigue endurance ratios at
Several studies have reported that the lamellar αþβ phases in an HIP
106 cycles of HIP samples with porosities of 33 vol% to 84 vol% were
sample are relatively more ductile or tougher than α0 -martensite in an
as high as 0.5, which is comparable to the endurance properties of
SLM specimen [12,23,24,27,29]. Moreover, fatigue crack blunting was
fully-dense Ti-6Al-4V and other Ti-based alloys.
found in an HIP sample in a previous study [24]. Thus, increasing the

8
M.-W. Wu et al. Materials Science & Engineering A 790 (2020) 139695

Data availability [15] S.J. Li, Q.S. Xu, Z. Wang, W.T. Hou, Y.L. Hao, R. Yang, L.E. Murr, Influence of cell
shape on mechanical properties of Ti-6Al-4V meshes fabricated by electron beam
melting method, Acta Biomater. 10 (2014) 4537–4547.
The raw data required to reproduce these results cannot be shared at [16] S.Y. Chen, J.C. Huang, C.T. Pan, C.H. Lin, T.L. Yang, Y.S. Huang, C.H. Ou, L.
this time as the data also forms part of an ongoing study. Y. Chen, D.Y. Lin, H.K. Lin, T.H. Li, J.S.C. Jang, C.C. Yang, Microstructure and
mechanical properties of open-cell porous Ti-6Al-4V fabricated by selective laser
melting, J. Alloys Compd. 713 (2017) 248–254.
Declaration of competing interest [17] S.Y. Choy, C.N. Sun, K.F. Leong, J. Wei, Compressive properties of functionally
graded lattice structures manufactured by selective laser melting, Mater. Des. 131
The authors declare no conflict of interest. (2017) 112–120.
[18] S.Y. Choy, C.N. Sun, K.F. Leong, J. Wei, Compressive properties of Ti-6Al-4V lattice
structures fabricated by selective laser melting: design, orientation and density,
CRediT authorship contribution statement Addit. Manuf. 16 (2017) 213–224.
[19] S.A. Yavari, S.M. Ahmadi, R. Wauthle, B. Pouran, J. Schrooten, H. Weinans, A.
A. Zadpoor, Relationship between unit cell type and porosity and the fatigue
Ming-Wei Wu: Conceptualization, Methodology, Data curation, behavior of selective laser melted meta-biomaterials, J. Mech. Behav. Biomed.
Supervision, Writing - original draft, Writing - review & editing. Jhewn- Mater. 43 (2015) 91–100.
Kuang Chen: Conceptualization, Methodology, Supervision, Funding [20] S. Zhao, S.J. Li, W.T. Hou, Y.L. Hao, R. Yang, R.D.K. Misra, The influence of cell
morphology on the compressive fatigue behavior of Ti-6Al-4V meshes fabricated
acquisition, Project administration. Bo-Huan Lin: Investigation, Visu­ by electron beam melting, J. Mech. Behav. Biomed. Mater. 59 (2016) 251–264.
alization. Po-Hsing Chiang: Investigation, Visualization. Mo-Kai Tsai: [21] S.J. Li, L.E. Murr, X.Y. Cheng, Z.B. Zhang, Y.L. Hao, R. Yang, F. Medina, R.
Investigation, Visualization. B. Wicker, Compression fatigue behavior of Ti-6Al-4V mesh arrays fabricated by
electron beam melting, Acta Mater. 60 (2012) 793–802.
[22] S.A. Yavari, R. Wauthle, J. van der Stok, A.C. Riemslag, M. Janssen, M. Mulier, J.
Acknowledgement P. Kruth, J. Schrooten, H. Weinans, A.A. Zadpoor, Fatigue behavior of porous
biomaterials manufactured using selective laser melting, Mater. Sci. Eng. C 33
The authors thank the Ministry of Science and Technology in Taiwan (2013) 4849–4858.
[23] B.V. Hooreweder, Y. Apers, K. Lietaert, J.P. Kruth, Improving the fatigue
for assistance under grant number MOST 104-2218-E-027-004 and performance of porous metallic biomaterials produced by selective laser melting,
MOST 107-2221-E-027-006. The authors also appreciate Atomic Craft Acta Biomater. 47 (2017) 193–202.
Corp., Taiwan, for preparing the SLM Ti-6Al-4V cellular samples used in [24] M.W. Wu, J.K. Chen, B.H. Lin, P.H. Chiang, Improved fatigue endurance ratio of
additive manufactured Ti-6Al-4V lattice by hot isostatic pressing, Mater. Des. 134
this study. (2017) 163–170.
[25] T. Vilaro, C. Colin, J.D. Bartout, As-fabricated and heat-treated microstructures of
References the Ti-6Al-4V alloy processed by selective laser melting, Metall. Mater. Trans. A 42
(2011) 3190–3199.
[26] Y.K. Kim, S.H. Park, Y.J. Kim, B. Almangour, K.A. Lee, Effect of stress relieving heat
[1] N. Mavros, T. Larimian, J. Esqivel, R.K. Gupta, R. Contieri, T. Borkar, Spark plasma
treatment on the microstructure and high-temperature compressive deformation
sintering of low modulus titanium-niobium-tantalum-zirconium (TNTZ) alloy for
behavior of Ti-6Al-4V alloy manufactured by selective laser melting, Metall. Mater.
biomedical applications, Mater. Des. 183 (2019) 108163.
Trans. A 49 (2018) 5763–5774.
[2] N. Soro, H. Attar, X. Wu, M.S. Dargusch, Investigation of the structure and
[27] R. Wauthle, B. Vrancken, B. Beynaerts, K. Jorissen, J. Schrooten, J.P. Kruth, J.
mechanical properties of additively manufactured Ti-6Al-4V biomedical scaffolds
V. Humbeeck, Effects of build orientation and heat treatment on the microstructure
designed with a Schwartz primitive unit-cell, Mater. Sci. Eng. A 745 (2019)
and mechanical properties of selective laser melted Ti6Al4V lattice structures,
195–202.
Addit. Manuf. 5 (2015) 77–84.
[3] P. Li, X. Ma, T. Tong, Y. Wang, Microstructural and mechanical properties of β-type
[28] P. Kumar, U. Ramamurty, Microstructural optimization through heat treatment for
Ti-Mo-Nb biomedical alloys with low elastic modulus, J. Alloys Compd. 815 (2020)
enhancing the fracture toughness and fatigue crack growth resistance of selective
152412.
laser melted Ti6Al4V alloy, Acta Mater. 169 (2019) 45–59.
[4] M.G. de Mello, C.A.F. Salvador, L. Fanton, R. Caram, High strength biomedical Ti-
[29] J.K. Chen, M.W. Wu, T.L. Cheng, P.H. Chiang, Continuous compression behaviors
13Mo-6Sn alloy: processing routes, microstructural evolution and mechanical
of selective laser melting Ti-6Al-4V alloy with cuboctahedron cellular structures,
behavior, Mater. Sci. Eng. A 764 (2019) 138190.
Mater. Sci. Eng. C 100 (2019) 781–788.
[5] L. Thijs, F. Verhaeghe, T. Craeghs, J.V. Humbeeck, J.P. Kruth, A study of the
[30] L. Huynh, J. Rotella, M.D. Sangid, Fatigue behavior of IN718 microtrusses
microstructural evolution during selective laser melting of Ti-6Al-4V, Acta Mater.
produced via additive manufacturing, Mater. Des. 105 (2016) 278–289.
58 (2010) 3303–3312.
[31] F. Brenne, T. Niendorf, Load distribution and damage evolution in bending and
[6] L.D. Bobbio, R.A. Otis, J.P. Borgonia, R.P. Dillon, A.A. Shapiro, Z.K. Liu, A.
stretch dominated Ti-6Al-4V cellular structures processed by selective laser
M. Beese, Additive manufacturing of a functionally graded material from Ti-6Al-4V
melting, Int. J. Fatig. 121 (2019) 219–228.
to Invar- Experimental characterization and thermodynamic calculations, Acta
[32] N. Vanderesse, A. Richter, N. Nu~ no, P. Bocher, Measurement of deformation
Mater. 127 (2017) 133–142.
heterogeneities in additive manufactured lattice materials by digital image
[7] W. Xu, M. Brandt, S. Sun, J. Elambasseril, Q. Liu, K. Latham, K. Xia, M. Qian,
correlation: strain maps analysis and reliability assessment, J. Mech. Behav.
Additive manufacturing of strong and ductile Ti-6Al-4V by selective laser melting
Biomed. Mater. 86 (2018) 397–408.
via in situ martensite decomposition, Acta Mater. 85 (2015) 74–84.
[33] P. K€ohnen, C. Haase, J. Bültmann, S. Ziegler, J.H. Schleifenbaum, W. Bleck,
[8] Y.J. Liu, H.L. Wang, S.J. Li, S.G. Wang, W.J. Wang, W.T. Hou, Y.L. Hao, R. Yang, L.
Mechanical properties and deformation behavior of additively manufactured
C. Zhang, Compressive and fatigue behavior of beta-type titanium porous
lattice structures of stainless steel, Mater. Des. 145 (2018) 205–217.
structures fabricated by electron beam melting, Acta Mater. 126 (2017) 58–66.
[34] Wu M.W., Chen J.K., Chiang P.H., Chang P.M., Tsai M.K., Compressive Property,
[9] H. Zhang, J. Zhao, J. Liu, H. Qin, Z. Ren, G.L. Doll, Y. Dong, C. Ye, The effects of
Deformation Behavior, and Fracture Mechanism of Selective Laser Melted Ti-6Al-
electrically-assisted ultrasonic nanocrystal surface modification on 3D-printed Ti-
4V Cellular Alloy with a New Cuboctahedron Structure, Submitted to Metallurgical
6Al-4V alloy, Addit. Manuf. 22 (2018) 60–68.
and Materials Transactions A (n.d.).
[10] S.A. Mantri, T. Torgerson, E. Ivanov, T.W. Scharf, R. Banerjee, Effect of boron
[35] A. Zargarian, M. Esfahanian, J. Kadkhodapour, S. Ziaei-Rad, Numerical simulation
addition on the mechanical wear resistance of additively manufactured biomedical
of the fatigue behavior of additive manufactured titanium porous lattice structures,
titanium alloy, Metall. Mater. Trans. A 49 (2018) 806–810.
Mater. Sci. Eng. C 60 (2016) 339–347.
[11] S. Ren, Y. Chen, T. Liu, X. Qu, Effect of build orientation on mechanical properties
[36] I. Sen, K. Gopinath, R. Datta, U. Ramamurty, Fatigue in Ti-6Al-4V-B alloys, Acta
and microstructure of Ti-6Al-4V manufactured by selective laser melting, Metall.
Mater. 58 (2010) 6799–6809.
Mater. Trans. A 50 (2019) 4388–4409.
[37] F. Li, J. Li, H. Kou, L. Zhou, Porous Ti6Al4V alloys with enhanced normalized
[12] M.W. Wu, P.H. Lai, The positive effect of hot isostatic pressing on improving the
fatigue strength for biomedical applications, Mater. Sci. Eng. C 60 (2016) 485–488.
anisotropies of bending and impact properties in selective laser melted Ti-6Al-4V
[38] V.F. Terentyev, S.V. Dobatkin, S.A. Nikulin, V.I. Kopylov, D.V. Prosvirnin, S.
alloy, Mater. Sci. Eng. A 658 (2016) 429–438.
O. Rogachev, I.O. Bannykh, Fatigue strength of submicrocrystalline Ti and Zr-2.5%
[13] S. Zhang, Q. Wei, L. Cheng, S. Li, Y. Shi, Effects of scan line spacing on pore
Nb alloy after equal channel angular pressing, Kovove Mater. 49 (2011) 65–73.
characteristics and mechanical properties of porous Ti6Al4V implants fabricated
[39] Y. Furuya, E. Takeuchi, Gigacycle fatigue properties of Ti-6Al-4V alloy under
by selective laser melting, Mater. Des. 63 (2014) 185–193.
tensile mean stress, Mater. Sci. Eng. A 598 (2014) 135–140.
[14] J. Kadkhodapour, H. Montazerian, A.C. Darabi, A.P. Anaraki, S.M. Ahmadi, A.
[40] Z.B. Zhang, Y.L. Hao, S.J. Li, R. Yang, Fatigue behavior of ultrafine-grained Ti-
A. Zadpoor, S. Schmauder, Failure mechanisms of additively manufactured porous
24Nb-4Zr-8Sn multifunctional biomedical titanium alloy, Mater. Sci. Eng. A 577
biomaterials- effects of porosity and type of unit cell, J. Mech. Behav. Biomed.
(2013) 225–233.
Mater. 50 (2015) 180–191.

You might also like