Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

Journal Pre-proof

Tensile and fatigue crack growth behavior of


commercially pure titanium produced by laser
powder bed fusion additive manufacturing

M. Tarik Hasib, Halsey E. Ostergaard, Qian Liu,


Xiaopeng Li, Jamie J. Kruzic

PII: S2214-8604(21)00192-5
DOI: https://doi.org/10.1016/j.addma.2021.102027
Reference: ADDMA102027

To appear in: Additive Manufacturing


Please cite this article as: M. Tarik Hasib, Halsey E. Ostergaard, Qian Liu,
Xiaopeng Li and Jamie J. Kruzic, Tensile and fatigue crack growth behavior of
commercially pure titanium produced by laser powder bed fusion additive
m a n u f a c t u r i n g , Additive Manufacturing, ()
doi:https://doi.org/10.1016/j.addma.2021.102027
This is a PDF file of an article that has undergone enhancements after acceptance,
such as the addition of a cover page and metadata, and formatting for readability,
but it is not yet the definitive version of record. This version will undergo
additional copyediting, typesetting and review before it is published in its final
form, but we are providing this version to give early visibility of the article.
Please note that, during the production process, errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
© Published by Elsevier.
Tensile and fatigue crack growth behavior of commercially pure titanium
produced by laser powder bed fusion additive manufacturing

M. Tarik Hasib, Halsey E. Ostergaard, Qian Liu, Xiaopeng Li, Jamie J. Kruzic

School of Mechanical and Manufacturing Engineering, The University of New South Wales
(UNSW Sydney), Sydney NSW 2052, Australia

Abstract
The effects of build orientation and post heat treatments on the tensile and fatigue crack

o f
growth (FCG) behavior of commercially pure titanium (CP-Ti) manufactured by laser

ro
powder bed fusion (LPBF) using grade 2 powder were examined. Two orthogonal build
orientations were used in conjunction with hot isostatic pressing (HIP) both above (950°C)

-p
and below (730°C) the β-transus temperature and property comparisons were also made to
commercially available wrought material. The HIP treatments coarsened the α grain
re
structure, reduced the tensile strength, and increased the fatigue crack growth threshold. The
LPBF materials were generally stronger and more fatigue resistant than the wrought material
lP

due higher interstitial oxygen and nitrogen content. Additionally, higher tensile strength
values were found for one build orientation with higher nitrogen content that was attributed
na

to the different thermal histories during LPBF. However, the build orientation effect was not
observed for the FCG behavior of the LPBF material and the FCG resistance at low growth
rates were mainly controlled by the grain size. This was in sharp contrast to the wrought
ur

material which showed strong anisotropy in the microstructure sensitive fatigue crack growth
Jo

regime due to strong crystallographic texture. At higher growth rates, FCG became
microstructure insensitive when the cyclic plastic zone size became of similar order of
magnitude to the grain size.

Keywords
Laser powder bed fusion; additive manufacturing; titanium; tensile strength, ductility, fatigue
crack growth; fatigue threshold

1. Introduction
Titanium and its alloys are widely used in the manufacturing of medical and dental implants
due to their high biocompatibility and corrosion resistance and relatively low elastic modulus
compared to other commonly used alloys such as stainless steel and CoCr [1, 2].
1
Commercially pure titanium (CP-Ti) is particularly attractive compared to other titanium
alloys due to its relatively higher biocompatibility, corrosion resistance, and osteointegration
ability [1]. Further benefits of CP-Ti include the absence of potentially hazardous alloying
elements and its low ion-formation tendency in aqueous environments. Additional
applications of CP-Ti include power generation, chemical/petrochemical production, heat
transfer, and aerospace due to its excellent corrosion resistance as well as good weldability
and fabricability [1]. Laser powder bed fusion (LPBF) is an additive manufacturing technique
which can fabricate complex metal parts by melting metal powder in successive layers using
a laser beam. Accordingly, LPBF fabrication of CP-Ti has received considerable recent
research attention [3-15]. Much of that research has focused on LPBF process parameter

f
optimization to achieve high density parts and evaluating the mechanical properties [5-15]. It

o
has been shown how microstructure can be controlled to achieve lath α, acicular α', and

ro
refined α' in LPBF manufactured CP-Ti parts by controlling the process parameters [9, 12].
Additionally, LPBF fabricated CP-Ti components tend to exhibit strong crystallographic
-p
texture along the build direction as a result of the directional heat flux and high thermal
gradient [13, 16]. However, it has been found that the crystallographic texture can be
re
weakened by controlling the laser power and the amount of preheating [13].
lP

Grain size, grain morphology, and crystallographic texture can greatly affect the mechanical
and physical properties of titanium and its alloys [17-42], and fatigue performance is a
na

property of particular importance for biomedical implants made from titanium [1]. In the case
of conventionally manufactured cast or wrought CP-Ti, there has been previous work devoted
to understanding how these microstructural factors affect the fatigue performance [32-41].
ur

Turner and Roberts found a trend to higher fatigue lives with finer grain sizes for CP-Ti as
the grain diameter was reduced from ~110 to 6 µm [39]. Over a similar grain size range (26 –
Jo

93 µm), negligible influence on the fatigue crack growth rates has been observed [32],
although Robinson and Beevers [34] found that increasing the grain size to >210 µm caused a
reduction in fatigue crack growth rates. This apparent discrepancy can be rationalized by
considering the majority of the fatigue life is spent initiating and growing small cracks. For
microstructurally small cracks, it has been found that coarse grains enable faster crack growth
which is attributed to fewer grain boundaries and crack deflections [38]. When considering
grain orientation, fatigue crack initiation in CP-Ti tends to occur mostly on the active basal or
prismatic slip planes [41], although under some test conditions crack initiation on twin planes
[41] or at grain boundaries [40] has been observed. Regarding fatigue crack growth, studies
of single crystal CP-Ti have shown that when the crystal is oriented for easy prismatic slip
(i.e., soft orientation), the crack will grow either by alternating shear on prismatic and
2
pyramidal slip systems or on two sets of pyramidal slip systems depending on the crack
growth direction [33, 36, 37]. In contrast, for loading along the c-axis (i.e., hard orientation)
cracks grow by deformation twinning which gives rise to much higher fatigue crack growth
resistance [33, 36, 37]. In this case, crack growth occurs by separation of the basal planes,
and in polycrystals this mechanism is observed when the loading axis is within 50° of the c-
axis of each grain [35]. Accordingly, fatigue life tends to improve when loading is along or
near the alpha phase c–axis of strongly textured titanium alloys since this orientation impedes
both crack initiation and propagation [17, 21, 30, 31, 37].

Despite the importance of fatigue performance for the targeted biomedical and dental
applications of additive manufactured CP-Ti, there are very few studies of fatigue properties

o f
in the literature [10, 43]. Santos et al. [10] found that the density of CP-Ti samples produced
by LPBF could be improved from ~92% to ~98% by optimizing the laser power and hatch

ro
spacing, and the improvement in density resulted in higher torsional fatigue strength [10].

-p
They further reported that hot isostatic pressing (HIP) gave nearly fully dense parts which
resulted in fatigue strengths that were comparable to wrought CP-Ti and adequate for dental
re
applications [10]. Okazaki and Ishino examined the fatigue life of CP-Ti produced by
selective laser sintering and found that fatigue lives comparable to dental cast or wrought CP-
lP

Ti could be achieved [43]. While such results suggest that LPBF is a promising
manufacturing route for targeted medical and dental applications, many questions remain
na

about how crystallographic texture may induce anisotropy in the fatigue properties, and how
fatigue performance is affected by various post heat treatments.
ur

To date, there have been no studies of the fatigue crack growth resistance of LPBF
manufactured CP-Ti. It is particularly important to understand the fatigue crack propagation
Jo

behavior because it allows the use of damage tolerant design methodologies in safety-critical
applications. In addition, discovering how LPBF produced microstructures and
crystallographic texture impact the fatigue crack propagation behavior of CP-Ti both with
and without post heat treatments will be essential to advancing this technology into real
world applications. To enhance our understanding of fatigue crack growth in CP-Ti, we used
compact tension C(T) samples manufactured in two orthogonal orientations to measure the
fatigue crack growth (FCG) resistance for LPBF produced CP-Ti. Two different post heat
treatments, HIP in the α field (730°C) and well above the beta transus temperature (950°C),
were applied to understand the influence of heat treatment on the FCG rates. Finally,
conventional tensile testing was performed for the same orthogonal build orientations to aid
in interpreting the trends in FCG behavior.

3
2. Materials and Methods

2.1 Powder and LPBF Process Parameters:

Commercially pure titanium powder (CP-Ti grade 2, TLS Technik GmbH, Germany) used in
this study was produced by argon gas atomization and had a spherical shape with ISO 13320
powder sizes of D10 = 19.19 µm, D50 = 35.70 µm, and D90 = 54.98 µm. A laser powder bed
fusion (LPBF) machine (SLM 125 HL, SLM Solutions Group AG, Germany) with a 1060 nm
wavelength Yb:YAG fiber laser (400 W max laser power; spot size of 100 µm) was used to
manufacture all samples in this study under a high-purity Ar atmosphere containing no more
than 0.02% oxygen on build plates preheated to 200°C before and during fabrication. LPBF

f
processing parameters for this powder were developed to obtain high relative density prior to

o
the manufacture of the mechanical test specimens. Energy density (E) is often considered to

ro
determine the quality and density of LPBF processed components. The energy density is
defined as [44, 45]:

E=
P
vht
,
-p (1)
re
where P is the laser power, v the scanning velocity, t the layer thickness, and h the hatch
spacing. Previous work on CP-Ti showed that choosing a fixed energy density of ~120 J/mm3
lP

produced high density CP-Ti parts [9]. Accordingly, four parameter sets were trialed with this
energy density to determine which achieved the best relative density. A summary of the
na

processing parameters is shown in Table 1. For each group, three cubes (10 mm × 10 mm ×
10 mm) were fabricated by using an initial 45º angle bidirectional laser scanning strategy
ur

with a 90° scan rotation between layers. The cubes were manufactured evenly spaced at the
middle location of the 120 mm x 120 mm build plate. Figure 1 shows the micrographs
Jo

illustrating the porosity achieved for each set of LPBF parameters. Relative densities
measured using Archimedes method [46] are shown in Table 1 and the process parameter set
named P4 (i.e., P = 165 W, v = 138 mm/s, t = 100 µm) was selected due to its > 98% relative
density with lower standard deviation and faster build time (i.e., a larger layer thickness)
compared to P3.

4
Table 1: The four sets of LPBF processing parameters that were examined and the
corresponding relative densities that were obtained for the 10 × 10 × 10 mm3 cubes.
Reference nominal density for CP-Ti was 4.51 gm/cm3 and relative density is given as
average ± standard deviation.

Relative
Group P (W) v (mm/s) t (µm) h (µm) E (J/mm3)
density
P1 165 460 30 100 119.5 96.3% ± 0.9
P2 200 555 30 100 120.2 97.1% ± 0.6
P3 250 695 30 100 119.9 98.2% ± 1.6
P4 165 138 100 100 119.5 98.2% ± 0.6

o f
ro
-p
re
lP
na
ur
Jo

5
P1 P2

P3 P4

o f
ro
-p
re
lP
na
ur

Figure 1: The presence of porosity for each set of LPBF parameters (see Table 1) is shown by
optical microscopy. Also shown is the 90º rotation bidirectional scanning strategy used for
Jo

each successive layer.


2.2 Test Samples:

Mechanical test samples were manufactured in two orthogonal orientations using the
processing parameters labeled P4 in Table 1 such that the build direction was either
perpendicular or parallel to the mechanical testing applied load (Figure 2). The nominal
dimensions of the tensile and compact tension C(T) specimens were selected in accordance
with the ASTM E8/E-8M [47] and ASTM E647 [48] standards. Figure 2 schematically shows
the orientations of the specimens on the build platform. Note that the full set of (N = 24)
tensile specimens and (N = 24) C(T) samples were fabricated on separate build platforms
while Figure 2 shows them on a single platform for brevity. For the tensile samples, the Z-

6
ASB tensile samples were placed on the left side of the build plate and the X-ASB tensile
samples occupied the rest of the area. For the C(T) samples, the C(T) coupons were
distributed evenly across the build plate. Fresh CP-Ti grade 2 powder was utilized to
manufacture all tensile samples. Unused power was sieved and recycled and added to fresh
powder for C(T) sample fabrication. This procedure was justified since impurity
concentration was found to be similar between the tensile and C(T) groups. Finally, for
comparison, a wrought plate of grade 2 CP-Ti was acquired from Prolog Titanium, Thailand,
to produce control samples. The 6.1 mm thick plate was heat treated at 700°C for 1 hr, stress
relieved at 540°C for 2 hr and then was air-cooled.

Twelve (N = 12) tensile coupons were built for each orientation with a total length of 77 mm,

o f
a gauge length of 12.5 mm, gauge width of 4.2 - 5.2 mm, and a thickness of 2.1 mm. The
naming convention used was (Z) for the build direction parallel to the applied load (X) for the

ro
build direction perpendicular to the applied load, which is shorthand for the ZX and XZ

-p
orientations, respectively, according to the ISO/ASTM 52921 standard [49]. For the wrought
CP-Ti plate, four (N=4) tensile samples were prepared by wire electro-discharge machining
re
(EDM) in each of two orientations. The loading axis was either oriented along the
longitudinal (rolling) direction (LD) or the transverse direction (TD).
lP

To investigate the fatigue crack growth properties in two orthogonal orientations, twelve (N =
12) compact tension C(T) coupon blanks were manufactured as shown in Figure 2. Also
na

shown in Figure 2, sample orientations for the C(T) samples were named by combining the
ISO/ASTM 52921 standard [49] terminology with that of ASTM E647 standard [48] for C(T
ur

samples. The first letter indicates the direction normal to the crack plane, and the second
letter indicates the crack growth direction (i.e., Z-X and X-Z). C(T) sample coupons were
Jo

produced with dimensions 32 mm × 31 mm to give a nominal C(T) sample width of W = 25.6


mm and a nominal thickness of B = 6.1 mm. The sample thickness satisfied the ASTM E647
W W
standard requirement of ≤ B ≤ 4 , noting that plane strain conditions are not required for
20

fatigue crack growth testing since they negligibly affect the results. Furthermore, sample
dimensions were sufficient to achieve the ASTM E647 standard small-scale yielding
requirement of (W-a) ≥ (4/𝜋)(𝐾max /𝜎YS )2 for all of reported fatigue crack growth results,
where a is the crack length, Kmax is the maximum stress intensity factor during the loading
cycle, and 𝜎YS is the yield strength. The C(T) specimen holes and notches were prepared by
wire EDM after LPBF fabrication. For the wrought CP-Ti plate, four (N=4) C(T) specimens
with the same nominal dimensions were prepared by water jet cutting for each of two
orientations, defined as T-L and L-T for wrought plate material according to Ref. [48]. All
7
tensile and C(T) sample surfaces were ground smooth using 220 grit SiC papers to eliminate
post-fabrication surface roughness.
Build direction

X-Z

o f
Z-X

ro
X

Z
Y
-p
re
lP

Figure 2: Tensile and C(T) sample orientations on the LPBF build plate. Note that the tensile
na

and C(T) samples were fabricated on separate build plates. Also, the EDM cut notches and
holes are shown here for reference only.
ur

2.3 Post Heat Treatments:


Jo

In this study, two different post heat treatments were employed. Hot isostatic pressing (HIP)
at 101 MPa pressure in an argon atmosphere was conducted at either a HIP temperature
(950°C) well above the beta transus temperature of 915°C or a HIP temperature (730°C) well
below the beta transus temperature. HIP was chosen to reduce porosity and defects while
coarsening the microstructure and the corresponding samples are were named HIP 950 and
HIP 730, respectively. The HIP soak time was 1 hr at each temperature while the heating rate
was 10°C/min and the cooling rate was ~3°C/min. A schematic summarizing the two post
heat treatments is presented in Figure 3.

8
60 min
950°C
Tβ=915ºC
Temperature (°C), pressure (MPa)

60 min
HIP 950
730°C
Furnace cooling

HIP 730
Furnace cooling

o f
ro
Time (min)
-p
Figure 3: Schematic illustrating the two HIP treatments selected for this study.
re
lP

2.4 Microstructure Analysis:

Prior to microstructure characterization, samples were prepared first by grinding using 300
na

grit SiC paper, then they were rough polished with a 0.9 µm diamond slurry, and lastly they
were fine polished using 0.04 µm colloidal silica mixed with 10% hydrogen peroxide. Phase
ur

identification was conducted using X-ray diffraction (XRD, X’pert MPD, Malvern
Panalytical, United Kingdom) with a Cu Kα-source and a scan rate of 0.02° s-1 from 2θ = 20°
Jo

to 100°. Grain size measurements were carried out using optical microscopy (OM, BX53,
Olympus, Japan) and scanning electron microscopy (SEM, JSM-7001F, JEOL, Japan) using
samples that were additionally etched by Kroll’s reagent (5% nitric acid, 10% hydrofluoric
acid, and 85% distilled water). Grain sizes were evaluated using the line intercept method
[50]. Crystallographic texture and kernel average misorientation (KAM) analyses was
performed by SEM at 20 kV using an electron backscatter diffraction (EBSD) detector and
EDAX OIM Analysis™ software (Version 5, AMETEK Inc, USA). The EBSD scanning step
size for the as-built (ASB) and HIP 730 samples was chosen to be 0.15 µm, while 0.4 µm
was selected for the HIP 950 samples due to its coarser grain size. The KAM was assessed up
to the third nearest neighbor kernel with a maximum misorientation of 5°. In order to
evaluate the chemical composition of the starting powder and the wrought and LPBF

9
samples, inductively coupled plasma atomic emission spectroscopy (ICP-AES) was used for
metallic elements and LECO combustion analysis was used for non-metallic elements. The
selected tensile sample size for both orientations was the half length of a fractured tensile
sample (i.e., 38 mm length). The C(T) sample used for chemical analysis was a half section
of a fractured specimen (i.e., 16 mm by 31 mm). The metal and carbon minimum detection
limit was 0.01 wt.% while for O, N, and H the minimum detection limit was 0.002 wt.%.

2.5 Mechanical Testing:

In all cases, mechanical tests were conducted for samples in the ASB condition as well as
after the HIP 730, and HIP 950 post heat treatments. Tensile tests were carried out in

f
displacement control (0.1 mm/min) in accordance with the ASTM E8/E-8M standard [47]

o
using a servohydraulic dynamic test system (Model 8852, Instron, USA) equipped with a 100

ro
kN calibrated load cell. For each group, four tensile samples were measured at room
temperature and a 12.5 mm gauge length extensometer (Model 2620-601, Instron, USA) was
-p
used to collect strain data. A two-way analysis of variances (ANOVA) was performed using
commercial software (SPSS, IBM Corporation, USA) to test the statistical significance of
re
each factor in affecting the results. Also, the statistical significance of pairwise differences
between groups was assessed using Tukey’s post-hoc test and p < 0.05 was considered
lP

statistically significant for all statistical tests.


na

Fatigue crack growth rates were measured using a servohydraulic dynamic test system
(Model 8872, Instron, USA) at room temperature with a 5 kN calibrated load cell by
following the ASTM E647 standard [48]. Test conditions employed were a sine wave
ur

frequency of 60 Hz and a load ratio of R = Pmin/Pmax = 0.1, where Pmin and Pmax were the
minimum and maximum loads applied during each loading cycle, respectively. To enable
Jo

observations of crack profiles, surfaces of the C(T) samples were mechanically polished
down to a 0.04 µm colloidal silica finish before starting each test using the procedure
described above for microstructure analysis. Crack lengths were continuously monitored via
a custom LabView program during each test using a backface strain gauge as per the ASTM
E647 standard [48]. Fatigue crack growth rate curves were constructed based on results from
four C(T) samples for each group.

Prior to testing, a precrack was extended about 2 mm from the machined notch by cyclic
loading at 60 Hz. Constant load amplitude testing was used to measure the Paris law growth
rates with an increasing stress intensity range, ΔK = Kmax – Kmin, where and Kmax and Kmin are

10
the stress intensity factors calculated from Pmax and Pmin, respectively. Plots of the fatigue
crack propagation rate, da/dN, versus ∆K were fit to the Paris law:

da
= C∆Km , (2)
dN

to determine the Paris law constants C and m. Decreasing ∆K tests following the ASTM E647
standard [48] were used to measure the fatigue thresholds, ∆Kth, using a constant normalized
K-gradient of -0.08 mm-1 until da/dN ≤ 10-10 m/cycle was reached.

Following testing, fractography of fatigue crack surfaces was performed by scanning

f
electronic microscopy (SEM, JSM-7001F, JEOL, Japan) and the fracture surface roughness

o
was measured using a laser scanning microscope (LSM, VK-X200, Keyence, Japan).

ro
Arithmetic average roughness, Ra, was determined in accordance with the JIS
B0601:2001(ISO/DIS 21920-2) standard [51]. Finally, crack profiles were observed on the
-p
C(T) samples surfaces by optical microscopy (OM, DSX510, Olympus, Japan)
re
The crack closure stress intensity, Kcl, was quantified using the ASTM compliance offset
method [48] as we have described in more detail in our previous work [52]. The effective
lP

stress intensity factor, ∆Keff, was calculated as:

∆Keff = Kmax - Kcl , (3)


na

for situations where Kcl > Kmin while ∆Keff = ∆K for situations where Kcl < Kmin.
ur

3. Results
Jo

3.1 Sample Density & Chemical Compositions

The relative density of P4 mechanical test samples are shown in Table 2 for the different
groups. The results shown in Table 2 indicate that HIP reduced the porosity to give higher
relative density and no cracking was identified by microscopy.

Table 3 gives the chemical composition (in wt.%) for the ASB and wrought samples.
Chemical analysis was provided by TLS Technik GmbH to ensure the CP-Ti powder was
within the grade 2 specifications (Table 3). It is noticeable that the oxygen weight % of LPBF
produced samples was consistently higher than the wrought sample, though it always
remained within the grade 2 CP-Ti specification. The weight % of nitrogen was also higher
for the LPBF produced samples compared to wrought CP-Ti; however, in this case it also

11
consistently exceeded the grade 2 CP-Ti specification (Table 3) making the final LPBF
produced sample better classified as grade 3.

Table 2: Relative density and microstructure characterization results for the various test
groups. All results are given as average ± standard deviation. Relative density data is given
for tensile samples and the reference density for CP-Ti was 4.51 gm/cm3

Sample ASB HIP 730 HIP 950

Relative Density (%) 96.5 ± 0.5 99.1 ± 0.9 99.4 ± 0.5

C(T) samples
Grain size/thickness 3.2 ± 1.0 4.6 ± 1.2 33.5 ± 6.3

o f
(µm)

ro
Tensile samples X-ASB Z-ASB X-HIP Z-HIP X-HIP Z-HIP
Grain size/thickness 730 730 950 950
(µm)

Wrought samples
3.5 ± 1.1 2.9 ± 0.8
-p
5.1 ± 1.6

Rolling/Longitudinal direction
4.2 ± 1.1 36.2 ± 11.2 29.5 ± 8.2

Transverse to the rolling direction


re
Grain size/thickness (LD) (TD)
(µm)
23.0 ± 6.2 32.5 ± 6.9
lP
na
ur
Jo

12
Table 3: Chemical composition specification for grade 2 CP-Ti (wt.%) along with the
measured compositions of the wrought plate and the LPBF produced tensile and C(T)
samples.
Identification Ti O Fe H C N
Grade 2
CP-Ti Bal. max.0.25 max. 0.30 max. 0.015 max. 0.08 max. 0.03
specification

Wrought
Bal. 0.17 0.05 <0.002 0.02 0.011
CP-Ti-2

Grade 2
CP-Ti Bal. 0.15 0.11 0.002 0.02 0.007
powder

o f
Tensile
Bal. 0.24 0.11 0.011 0.02 0.031
X-ASB

ro
Tensile
Bal. 0.23 0.11 0.004 0.02 0.051
Z-ASB -p
re
C(T)
Bal. 0.23 0.11 0.006 0.03 0.041
X-Z ASB
lP

C(T)
Bal. 0.19 0.12 0.008 0.02 0.048
Z-X ASB

3.2 Microstructure & Texture:


na

The XRD diffraction patterns (Figure 4e) indicated the existence of an HCP phase in all
samples, though it was not possible to distinguish martensitic α' from the α phase by XRD
ur

due to their similar HCP structures as is commonly reported in the literature. The
Jo

microstructure of ASB samples (Figure 4a) comprised of a large number of elongated


acicular grains and a small number of equiaxed grains, similar to that observed by others [5,
12, 13]. Based on this observation, it is presumed that the ASB sample contained mostly
acicular α' martensite due to fast cooling below the β-transus temperature (Tβ ~ 915 °C for
CP-Ti grade 2). The average α'/α grain thickness of the ASB samples was around 3.2 µm
(Figure 4a,5a and Table 2) and the variability between the X and Z tensile samples was
minor. Thus, the processing parameters were considered to be quite robust for different print
geometries. Heat treatment below the β-transus temperature (HIP 720) was presumed to fully
convert the structure to the α phase and resulted in a coarser average α grain thickness of
around 4.6 µm while changing the morphology to include more equiaxed grains (Figure
4b,5b and Table 2). Heat treatment above the β-transus temperature (HIP 950) significantly

13
coarsened the microstructure giving elongated blocky grains along with some equiaxed grains
and the average α grain thickness increased to around 33.5 µm (Figure 4c,5c and Table 2).
Overall, when considering the X and Z groups separately, the X oriented samples of all
conditions were slightly coarser than the Z oriented samples (Table 2). Finally, the wrought
CP-Ti had a recrystallized, equiaxed grain structure (Figure 4d and Figure 5d,e) and the
average grain size was measured on two orthogonal planes. For the plane with the LD
direction oriented perpendicular to the plane (Figure 5d), the grain size was around 23.0 µm
(Table 2). For the plane with the TD direction oriented normal to the plane (Figure 5e), the
grain size was around 32.5 µm (Table 2).

o f
ro
-p
re
lP
na
ur
Jo

14
(a) (b)

(c) (d)

o f
ro
50 µm

(e)
-p = hcp
re
lP
Intensity (a.u.)

HIP 950
na

ASB
ur
Jo

Wrought

Position (angle

Figure 4: Optical micrographs and XRD results for the various CP-Ti sample groups. Optical
micrographs are shown for (a) ASB, (b) HIP 730, (c) HIP 950 and (d) wrought CP-Ti
samples. For panels (a) – (c) the LPBF build direction is orientated out of the page and for
panel (d) the rolling direction is oriented out of the page. Panel (e) displays the XRD plots of
ASB, HIP 950 and wrought CP-Ti samples. XRD results for HIP 730 were similar to ASB
and HIP 950 and are not shown.

15
(a) (b) (c)

f
(d) (e)

o
ro
-p α phase
re
lP

70 µm
na

Figure 5: EBSD results showing inverse pole figure maps for each microstructure: (a) ASB,
(b) HIP 730, (c) HIP 950, (d) wrought LD and (e) wrought TD. For panels (a) – (c), the LPBF
ur

build direction (Z) and loading direction are both oriented out of the page. The black color
represents zero EBSD scan index. For panel (d) the rolling direction and panel (e) transverse
direction is oriented out of the page. Black arrows point out equiaxed grains in panels (b) &
Jo

(c).
EBSD texture results are presented in Figure 6 through Figure 8 as pole figures. Of the 12
possible Burger’s relations, (110)β // (0001)α and <111>β // <1120>α are preferred [1, 13,
53], although texture generated due to other variant selection has been also reported [54-56].
For LPBF manufactured titanium alloys it has generally been found that the α/α' laths tend to
follow the preferred orientation with their c-axes at roughly 30-60° relative to the build
direction [57-59]. In the present study, the (0001) pole figure of Z-ASB (Figure 6a) sample
shows a tendency for the α/α' laths to be oriented with their c-axes at ~30-45° to the build
direction. Other variants were observed as well with lower frequency, for example, some α/α'
laths were orientated with their c-axes at ~90° to the build direction. Similar texture with the
α/α' phase c-axis orientation at ~30-45° to the loading direction can be seen for X-ASB

16
samples (Figure 7a); however, in this case some grains are orientated with their c-axis
parallel (0°) to the loading direction. It is important to note that for all X samples the loading
and build directions were perpendicular to each other while for Z samples the loading and
build directions were parallel. HIP 730 samples for both Z and X orientation show (Figure 6b
and Figure 7b) some similar features as the ASB samples since HIP below Tβ does not
change the prior β grains and some of the acicular α structure remains intact. Figure 6c and
Figure 7c represent (0001) pole figures of HIP 950 samples of Z and X orientation,
respectively. Equiaxed α grains for both X & Z oriented HIP 950 samples were mostly
orientated with the c-axis at a 40-45° angle to tensile loading direction, while for the X
orientation there was also a peak oriented at ~0°. Finally, the wrought plate generally showed

f
more texture than the LPBF produced material. The LD orientated wrought CP-Ti samples

o
(Figure 8(a)) had the (0001) basal poles mostly concentrated ~40-70° away from the loading

ro
direction toward the TD (transverse direction). In contrast, the TD orientated samples had the
(0001) basal poles mostly concentrated ~80-90° from the loading direction.

Z-ASB
-p
Z-HIP 730
re
lP
na

(a) (b)
ur

Z-HIP 950
Jo

BD,FD

(c)

Figure 6: (0001) pole figures for Z orientated ASB and HIP samples. (a) Z-ASB, (b) Z-HIP
730, and (c) Z-HIP 950. BD= build direction, FD= loading direction. Here both BD and FD
are out of page and parallel to each other.

17
X-ASB X-HIP 730

(a) (b)

X-HIP 950
BD

o f
FD

ro
-p
re
(c)
lP

Figure 7: (0001) pole figures for X orientated ASB and HIP samples: (a) X-ASB, (b) X-HIP
730, and (c) X-HIP 950. BD= build direction, FD= loading direction, Here FD is out of page
while BD is perpendicular to FD.
na

LD TD
ur
Jo

(a) (b)

ND ND

FD,LD FD,TD

Figure 8: (0001) pole figures for the (a) LD and (b) TD oriented wrought samples. LD =
longitudinal/rolling direction, FD = loading direction, ND = normal direction to rolling and
TD = transverse direction to rolling. Here FD is out of the page.

18
Kernel average misorientation (KAM) maps and average KAM values are displayed in
Figure 9 for the various sample types. The decline in mean KAM value after HIP, especially
for the HIP 950 treatment, suggests a decrease in dislocation density [60].

X- ASB X- HIP 730 X - HIP 950


(a) (b) (c)

o f
ro
0.75 0.63 0.21 5°

(d)
Z- ASB
(e)
Z- HIP 730
-p (f)
Z- HIP 950 0°
re
lP
na

70 μm
ur

0.82 0.66 0.32


Jo

Figure 9: Kernel average misorientation (KAM) maps with the mean KAM value given under
each map. (a) X-ASB, (b) X-HIP 730, (c) X-HIP 950, (d) Z-ASB, (e) Z-HIP 730, and (f) Z-
HIP 950. The color scale ranges from blue to red for 0° to 5° of misorientation whereas black
lines indicate grain boundaries. All cross sections were taken perpendicular to the loading
direction.

3.3 Mechanical Properties:

Table 4 summarizes 0.2% offset yield strength, 𝜎YS , and ultimate tensile strength, 𝜎UTS ,
results. Build orientation and heat treatment significantly affected the 𝜎YS and 𝜎UTS , with no
significant interaction, according to the two-way ANOVA results. Both 𝜎YS and 𝜎UTS
decreased monotonically with increasing heat treatment temperature and α grain size, with

19
the Z orientation being consistently stronger. The Z orientation ASB samples had the highest
average yield strength value of 630 MPa and average 𝜎UTS of 720 MPa, although the lower
values for the HIP-730 samples were not statistically significant for either orientation (Table
4). Also, the Z-ASB sample yield strength was higher compared to some previous studies that
used the same build orientation [9, 11, 61]. Finally, the wrought CP-Ti was consistently
weaker than the LPBF material and had lower yield strength when tested in the longitudinal
(rolling) direction (LD) versus the transverse direction (TD), although this trend was reversed
for the 𝜎UTS (Table 4). The TD wrought samples were the most ductile (~31.5%) which was
closely followed by the LD wrought samples (~29%). The LPBF produced samples were
generally less ductile (6.3 to 15.1%) than the wrought samples, which can be attributed to

f
their much higher strengths. Although the X orientation of the LPBF samples had

o
consistently lower strength, it did not have consistently higher ductility and the Z-HIP 950

ro
samples exhibited the best ductility of all of the LPBF produced samples.

-p
Table 4: Summary of the tensile test results presented as mean ± standard deviation.

Ultimate Tensile %
re
Yield Strength* %
Sample Strength* 𝝈𝐔𝐓𝐒 Reduction
𝝈𝐘𝐒 (MPa) Elongation*
(MPa) in area*
lP

X 521 ± 13.1a 607 ± 16.5g 10.4 ± 2.6m 10.0 ± 3.3p


ASB
Z 630 ± 20.6b 720 ± 22.5h 8.3 ± 1.6m 7.2 ± 2.3p
na

HIP X 512 ± 14.3a 587 ± 21.6g 7.3 ± 1.3m 9.5 ± 2.4p


730 Z 622 ± 10.1b 716 ± 12.6h 15.1 ± 3.1n 11.2 ± 1.6p
HIP X 482 ± 12.7c 573 ± 26.6i 6.3 ± 1.3m 7.7 ± 0.8p
ur

950 Z 573 ± 33.3d 662 ± 38.8j 7.4 ± 2.2m 6.6 ± 2.9p


Jo

LD 317 ± 6.8e 481 ± 11.9k 28.9 ± 1.2o 29.2 ± 1.4q


Wrought
TD 382 ± 15.4f 457 ± 13.7l 31.6 ± 1.5o 31.4 ± 1.6q
*
Identical superscripts in an individual column are used to indicate no statistically significant
difference between the mean values.
Fatigue crack growth rate results are shown in Figure 10 while Table 5 summarizes the
threshold values deduced from the plots. The fatigue thresholds demonstrated an inverse
relationship to strength for the LPBF produced material, with ∆Kth rising monotonically with
increasing heat treatment temperature and α grain size. However, the wrought material did
not follow the same trend as the LPBF material; rather, it had the lowest strength but not the
highest ∆Kth. Furthermore, the lowest fatigue threshold of all cases was found for the T-L
orientation of the wrought material, which also had the lowest yield strength. While the ∆Kth

20
values for the LPBF material showed little effect of sample orientation, there was significant
anisotropy to the wrought material with ∆Kth for the L-T orientation being approximately
56% higher than that of the T-L orientation.

At higher growth rates, Figure 10 reveals that the fatigue crack growth rate curves displayed
a bilinear shape, which is a commonly observed feature for (α+β) titanium alloys (e.g., Ti-
6A-4V) with basketweave or lamellar microstructures [23-28, 62]. Consequently, the
transition stress intensity range, ∆KT, where the Paris law constants change was deduced by
fitting Eq. (2) both above and below the transition as described in our previous work [52]. All
of the deduced values are given in Table 5. Similar to (α+β) alloys [23-28, 52, 62], the
hypertransitional exponents, mA, were relatively insensitive to microstructure across all

o f
materials. In contrast, the hypotransitional exponents, mB, for the LPBF produced material

ro
increased with coarsening of the microstructure (Table 5), has been observed for LPBF
produced Ti-6Al-4V [52]. However, like with the fatigue threshold, the wrought material did

-p
not follow the same trend with the mB values being most similar to the very fine grained ASB
material. In addition, increasing heat treatment temperature and α grain size gave improved
re
resistance to fatigue crack growth in the hypotransitional Paris region (Figure 10), which can
be quantified by the decreasing Paris coefficients, CB seen in Table 5.
lP
na
ur
Jo

21
Table 5: Fatigue crack growth rate, crack closure, and fracture surface roughness results.

Paris Law Constants


C units: (m/cycle)/(MPa√
m
∆Kth m) Transitional Kcl ∆Keff,th
Ra
Sample (MPa Hypo- Hyper- ΔKT (MPa (MPa (MPa Kcl/Kmax
(μm)
√m) transitional transitional √m) √m) √m)
Region Region
mB CB mA CA
-
6.3 ×10 2.8 ×
ASB X-Z 2.8 3.8 12 2.3 -10 11.7 2.8 10
10
-

f
3 × 10 3.5 ×
ASB Z-X 2.9 4.2 3.2 11.5 2.9 13

o
12 -11
10
HIP 730 4.3 × 1.4 ×

ro
3.3 4.7 -13 2.6 -10 14.7 0.75 2.9 0.20 16
X-Z 10 10
-
HIP 730 8.9 × 1 × 10
Z-X
HIP 950
3.2

4.1
4.5

5.1
10
-13

1.6 ×
-13
2.8

2.5
10

1.7 ×
-10
-p 14.6

15.1
0.68

0.94
2.8

3.6
0.19

0.20
18

20
re
X-Z 10 10
HIP 950 2.8 × 4.1 ×
3.9 4.7 -13 2.9 -11 16.3 0.89 3.4 0.21 27
lP

Z-X 10 10
7.9 × 6.8 ×
T-L 1.8 3.9 -12 2.9 -11 10.6 0.53 1.4 0.25 15
10 10
na

1.8 × 3.1 ×
L-T 2.8 4.3 -12 3.1 -11 11.5 0.72 2.4 0.23 18
10 10
ur
Jo

22
of
ro
-p
re
a lP
u rn

Figure 10: Plots of the fatigue crack growth rates, da/dN, versus the applied stress intensity
Jo

range, ∆K. Load ratio was R=0.1 and test frequency was 60 Hz. The different build
orientations are distinguished by different symbols while the different colors reflect the
various post heat treatments used in this study: red = ASB, green = HIP 730, blue = HIP 950,
black = wrought.

23
mA,CA ∆KT mB,CB

ASB

HIP
730

HIP

of
950

ro
L-T
-p
re
T-L
lP

100 µm
a

-6 -7
Paris region, ~10 - 10 m/cycle
rn

Crack growth direction


u

Figure 11: Crack profiles in the near ∆KT transition regime for X oriented LPBF samples and
Jo

for both T-L and L-T wrought samples. The black dash-dot line roughly indicates where the
∆KT transition occurred in the bilinear Paris region. The right of the dash-dot line represents
the hypotransitional region while the left of the dash-dot line represents hypertransitional
region.

Figure 11 shows fatigue crack profiles for growth in the Paris regime near the ∆KT transition.
Although some divergence of the cracks from the C(T) sample symmetry plane is seen in
Figure 11, the majority of samples deviated less than 10° which was considered to be
insignificant. Four cases (ASB Z; HIP 730 X & Z, and HIP 950 X) saw cracks that diverged
between 10° and 20° from the plane of symmetry; however, the effect on the fatigue crack
growth rates is expected to be negligible since deviations under 20° are allowed by the

24
ASTM E647 standard [48]. Finally, roughness data measured near the fatigue threshold
region of the fracture surfaces (Table 5) indicated that larger α grain size tends to induce, on
average, rougher crack paths among the LPBF samples.

Deduced values of the crack closure stress intensity factor, Kcl, and the effective stress
intensity range, ∆Keff,th, determined at the fatigue threshold are given in Table 5. Linear
unloading compliance down to Kmin for the ASB samples suggested a closure free crack for
those samples and an effective fatigue threshold equal to ∆Kth. For all other samples, the
values of Kcl were similar and fell between 0.5 – 0.9 MPa√m. While the closure free
thresholds, ∆Keff,th, were identical for the ASB and HIP 730 with similar fine acicular grain
structures, the ∆Keff,th values were consistently higher for the HIP 950 samples and lower for

of
the wrought samples. This suggests that microstructure morphology differences were more
important in controlling the thresholds than differences in crack closure or the α grain size.

ro
4. Discussion:
-p
re
4.1 Phase formation and microstructure:
lP

The microstructure of the ASB samples comprised of a large number of elongated acicular
grains and a small number of equiaxed grains, similar to that observed by others [5, 12, 13].
a

However, the α grain thickness of the X-ASB build orientation was coarser than the Z-ASB
rn

build orientation, and this trend was maintained for the heat treated samples as well (Table 2).
Since both X and Z orientated tensile samples were fabricated on the same build platform,
u

this difference was attributed to the larger layer area and longer layer scan time for the X
Jo

build orientation which gave more heat input and time for grain coarsening. After the HIP
730 treatment (below β transus), the α grain size became 44% thicker than for the ASB
condition and the equiaxed grains increased in number; however, much of the acicular
structure of the ASB samples remained. In contrast, heat treatment above the β-transus (HIP
950) and slow furnace cooling completely destroyed the acicular structure resulting in large
blocky grains along with some equiaxed grains. However, the X-HIP 950 build orientation
still maintained a coarser microstructure than the Z-HIP 950 build orientation. This is
attributed to the presence of ~0.1% Fe (Table 3) that stabilizes small amounts of the β phase,
which will resist coarsening in the α grain structure [1]. In this regard, the coarser initial β
grain size of the X-HIP 950 samples will lead to a coarser retained β phase that puts less
constraint on the coarsening the α grain structure compared to the Z-HIP 950 samples.

25
4.2 Microstructure and Build Orientation Effects on Strength:

Post heat treatment of the LPBF produced samples resulted in lower yield and tensile strength
(Table 4); however, this difference was only statistically significant for the HIP 950
treatment. Additionally, HIP coarsened the grain size with increasing temperature (Figure 5)
and decreasing strength with microstructure coarsening is commonly observed for CP-Ti
alloys [1, 63-65]. Furthermore, Figure 9 displays decreasing KAM values which suggest
there is a reduction of dislocation density after heat treatment. However, the softening effect
with heat treatment is weak, which may be attributed to the fact that the main strengthening
mechanism for CP-Ti is solid solution strengthening [1, 66-72].

of
Based on Table 3, the carbon constant was relatively low to induce much solid solution
strengthening [66] and was also constant for wrought and LPBF samples that exhibited very

ro
different strengths. In contrast, the oxygen and nitrogen pickup during LPBF was sufficient to
cause significant strengthening [66] and the higher content of these elements in the LPBF
-p
samples explains the higher strength relative to the wrought alloy. It is also interesting to note
that while the oxygen content stayed within the grade 2 CP-Ti specification after the LPBF
re
process, the nitrogen content generally exceeded the specification. Oxygen and nitrogen are
lP

the most common expected impurities in the LPBF environment, and nitrogen in particular is
a potent solid solution strengthener of α titanium in small amounts less than 0.1 wt.% due to
its strong influence on the c/a ratio in the HCP lattice [66, 73]. Based on the present results,
a

small amounts of nitrogen impurity pickup from the LPBF chamber appear to have a more
rn

significant effect on the final properties compared to oxygen.


u

The σYS and σUTS values for all LPBF samples showed the Z build orientation was
Jo

consistently stronger than the X orientation. Considering the effect of grain size on strength
was relatively weak, the strength differences could not be fully attributed to the slight grain
size differences seen in Table 2. While texture is another possible explanation, the ASB
texture shown in Figure 6a and Figure 7a indicate that for both Z and X oriented samples
there are a large amount of α grains oriented with their c-axis ~30-45° from the loading
direction. Since both orientations have a similarly large proportion of grains oriented for easy
⃗ basal slip, texture does not easily explain the large strength differences. Instead of texture
𝒂
or grain size effects, much of the strength differences observed for the various samples in this
study are attributed to differences in the chemical compositions (Table 3).

While the oxygen content was roughly constant across all LPBF samples, the nitrogen
content was constantly higher for the Z oriented samples (Table 3). A recent study has shown
26
that nitrogen content tends to increase with increasing laser power during LPBF, and this was
attributed to higher temperatures and longer times at temperature associated with the higher
laser power [74]. In the present work, the laser power was constant for all tensile samples;
however, the gauge sections of the Z oriented samples were built after the X oriented samples
were completed on the build plate (Figure 2). Thus, the Z oriented samples had less time to
cool between each build layer and likely experienced higher temperatures leading to more
nitrogen pick up. While oxygen impurity levels also increased during fabrication, they were
similar for the different build orientations (Table 3). To quantify the role of nitrogen, Table 6
shows the predicted solid solution strengthening based on Labusch’s model of solid solution
strengthening [75]. The mathematical expression of Labusch model for the increase of the

of
0.2% yield strength due to solid solution nitrogen, ∆𝜎YS,N , may be written as [76]:

1⁄3
𝑐 2⁄3

ro
𝐹4
𝑚𝑤
∆𝜎YS,N = (4𝐺𝑏 9)
, (4)
𝑆𝐹

where c is the solute nitrogen atomic fraction, SF is the Schmid factor, G is the shear
-p
modulus, b is the magnitude of the Burger’s vector, w represents the area affected by the
interaction of the solute with an edge dislocation and Fm is the maximum interaction force of
re
1⁄3
𝑚𝑤 𝐹4
Ti and N. The value of (4𝐺𝑏 9) may be obtained by plotting the increment of 0.2% ∆σYS
lP

𝑐 2⁄3
against the material’s factor which follows a linear relationship through the origin [76,
𝑆𝐹

77]. Table 6 compares the expected increment of 0.2% ∆σYS from Eq. (4) using the value of
a

1⁄3
𝐹4
rn

𝑚𝑤
(4𝐺𝑏 9) = 3,293 taken from [76, 77] with the experimental 0.2% ∆σYS variations between

the two orthogonal build orientations for each group due to nitrogen solid solution
u

strengthening. The results in Table 6 show that the majority of the strength difference can be
Jo

attributed to the nitrogen difference. The remainder of the difference is attributed to the
aggregate of minor effects due to variations in grain size, texture, dislocation density, residual
stress, etc.

27
Table 6: Calculated nitrogen solid solution strengthening effect compared to the experimental
results.

Sample LPBF Schmid (0.2%) Yield Experimental Calculated


nitrogen factor stress, σYS Yield Stress Yield Stress
(at. %) from (MPa) Increment, Increment,
EBSD ∆σYS,N (MPa) ∆σYS,N (MPa)
X-ASB 0.104 0.401 521
109 58
Z-ASB 0.173 0.332 630
X-HIP 730 0.104 0.408 512
110 58
Z-HIP 730 0.173 0.336 622
X-HIP 950 0.104 0.434 482
91 52
Z-HIP 950 0.173 0.365 573

of
ro
In contrast to the LPBF samples, the differently oriented wrought samples all came from the
same highly textured plate with the same chemical composition, suggesting that
-p
crystallographic texture was responsible for the anisotropy in the tensile properties. As should
be expected, the longitudinal oriented samples (LD) exhibited strong basal/transverse texture
re
(Figure 8a) and lower σYS values compared to the transverse oriented samples (TD) that
lP

exhibited strong transverse texture (Figure 8b) [1, 65, 78]. For the latter case, the c-axis of the
α grains are mostly angled at ~90° to loading direction. As a result, deformation is more
difficult as the pyramidal slip system (⃗⃗⃗𝒂 +⃗⃗𝒄 ) requires higher stress for activation compared
a

to basal or prismatic slip.


rn

4.3 Near-Threshold Fatigue Crack Growth:


u

The fatigue thresholds, ∆Kth, for the LPBF produced CP-Ti increased with increasing heat
Jo

treatment temperature and α grain size (Figure 10 and Table 5). Additionally, roughness
measurements (Table 5) suggest that the crack path becomes more torturous after heat
treatment. The increased roughness is thought to be responsible for the crack closure that
develops with heat treatment (Table 5) by the mechanism of roughness induced crack closure
[79-84]. Comparing the ASB and HIP 730 samples, once the difference in crack closure is
taken into account, the effective fatigue thresholds, ∆Keff,th, are identical. Since these samples
had similar α grain sizes, the suppression of crack closure for the ASB samples was likely
related to residual stresses that were relieved by heat treatment. Comparing the HIP 730, HIP
950, and wrought samples, the contribution of crack closure to the fatigue threshold was
similar in all cases, suggesting that other microstructural factors were responsible for the
differences in their fatigue thresholds.
28
Rather, the trend of increasing fatigue threshold with increasing α grain size is better
explained by the blocked slip band model whereby continued fatigue crack growth requires
effective slip transfer from one grain to the next [85-88]. For finer grained materials,
dislocation pile-ups at grain boundaries occur more readily and provide ample stress to
activate deformation in the next grain. Consequently, finer grained materials have lower
fatigue thresholds. A quantitative description of this model reveals that the fatigue threshold
should be proportional to the characteristic microstructural dimension, d, and the yield stress,
𝜎YS , according to [85, 86]:

∆𝐾th ∝ 𝜎YS √𝑑. (5)


The α grain size determines the effective slip length in CP-Ti and should be taken as the

of
characteristic microstructural dimension [1]. A plot of ∆𝐾th versus 𝜎YS √𝑑 appears is nearly

ro
linear for the three LPBF groups (Figure 12a), and the linear relationship improves once the
effect of crack closure is removed and Eq. (5) is replotted in terms of ∆Keff,th (Figure 12b).

-p
Improvement of the linear fit in Figure 12b is attributed to the effective fatigue threshold
being more directly related to slip transference across the grain boundaries. Thus, the higher
re
threshold of HIP 950 is controlled by the change in slip behavior due to microstructure
coarsening rather than due to changes in crack closure.
a lP
u rn
Jo

29
6
(a)
HIP 950
HIP 730
∆Kth (MPa√m)
4

2
ASB Wrought

0
0 1 2 3 4
σYS√d (MPa√m)

of
6

ro
(b)
HIP 950
∆Keff,th (MPa√m)

HIP 730
4
-p
re
2
ASB Wrought
lP

0
0 1 2 3 4
a

σYS√d (MPa√m)
rn

Figure 12: (a) Correlation between the microstructure size scale plotted as 𝜎YS √𝑑 (where d is
defined as the α grain size from Table 2), and the fatigue threshold, ∆Kth, according to Eq. 5.
u

(b) shows the same plot in terms of the effective fatigue threshold, ∆Keff,th, instead of ∆Kth. In
Jo

both (a) & (b), the line indicates the linear fit for the three LPBF microstructures (ASB, HIP
730, & HIP 950).
In both Figure 12a and Figure 12b, the wrought CP-Ti microstructure sits far from the linear
relationship. Furthermore, the wrought samples machined in the orthogonal L-T and T-L
orientations (Table 5 and Figure 10) exhibited significantly more anisotropy in ∆Kth
compared to the LPBF samples built in the orthogonal Z-X and X-Z orientations (Figure 2).
While the Z-X and X-Z LPBF orientations had essentially identical thresholds, the wrought
L-T orientation was found to have a fatigue threshold ~ 56% higher than the T-L orientation.
The higher threshold for the L-T orientation is attributed to the strong texture and difficulty
of slip or twining. Figure 8a shows the (0001) pole figure for the L-T samples indicating that
many grains are oriented with the c-axis nearly aligned with the loading direction. Studies on
fatigue crack growth in single crystals of titanium oriented with loading along the c-axis have
30
found that the unfavorable slip plane orientation instead favors crack growth by deformation
twinning and separation of the basal planes, giving higher fatigue crack growth resistance
[33, 36, 37]. Further evidence for this crack growth mechanism is seen in Figure 13 where a
faceted mode of fatigue crack propagation is seen for the L-T samples. Thus, the presence of
“hard” oriented grains appears to be responsible for the elevated fatigue threshold relative to
the T-L orientation.

In contrast, the c-axes for the α grains of the T-L samples were mostly orientated at ~80-90°
angle (Figure 8b) to the fatigue loading direction whereby the crack should grow either by
alternating shear on prismatic and pyramidal slip systems or on two sets of pyramidal slip
systems depending on the crack growth direction [33, 36, 37]. Alternating slip gives rise to

of
easier crack growth and fatigue striation formation in the microstructure sensitive crack
growth regime and Ward-Close et al. have reported one-to-one correlation of fatigue

ro
striations and fatigue cycles for CP-Ti down to ~10-9 m/cycle [35]. At slower growth rates

-p
similar striation features are still observed, though the one-to-one correspondence is lost,
most likely due to non-uniform growth along the crack front. Figure 14 shows the fracture
re
surface for the T-L orientation where a clear striation morphology is observed in contrast to
the faceted features seen in Figure 13. Thus, the large difference anisotropy in fatigue crack
lP

growth response can be explained by the crystallographic texture.

Overall, the fatigue fracture surfaces of the LPBF samples showed an absence of the faceted
a

separation of the basal planes seen for the L-T samples, with representative micrographs
rn

shown in Figure 15 for the ASB condition. The HIP 950 LPBF samples were similar in grain
size to the wrought material and demonstrated consistently higher fatigue crack growth
u

resistance in the microstructure sensitive regime, which is attributed to the higher strength
Jo

caused by the higher oxygen and nitrogen content that gives higher resistance to striation
formation. A trend of slower fatigue crack growth with increasing interstitial concentration
and strength has been reported previously by Robinson and Beevers for wrought CP-Ti [32].
Overall, while the higher fatigue resistance observed for the LPBF samples could likely be
achieved in wrought material with similar interstitial content, the lack of anisotropy in the
fatigue crack growth properties suggests potentially superior performance for LPBF produced
components in real world applications relative to components machined from commercially
available wrought material.

31
(a) (b)

10 µm

(c) (d)

of
ro
1 µm

Crack direction
-p
re
Figure 13: L-T sample fatigue fracture surface in the near-threshold region. (a) shows the
lP

overall transgranular mode of crack propagation while (b) - (d) show higher magnification of
the faceted steps with little plastic deformation. The approximate ∆K = ~2.7 MPa√m for all
figures. Scale bar is identical for (b) – (d) and is shown in panel (d).
a
u rn
Jo

32
(a) (b)

10 µm 1 µm
(c)

of
ro
1 µm
-p
Crack direction
re
Figure 14: T-L sample fatigue fracture surface in the near-threshold region. (a) shows the
lP

overall transgranular mode of crack propagation while (b) & (c) show higher magnification
images where sub-micrometer striations can be identified (arrows). The approximate ∆K =
~1.7 MPa√m for all figures.
a
u rn
Jo

33
(a) (b)

10 µm 1 µm

(c)

of
ro
1 µm

Crack direction
-p
re
Figure 15: Fatigue fracture surface for an ASB sample in the X-Z orientation near the fatigue
lP

threshold region. (a) shows the overall transgranular mode of crack propagation while (b) &
(c) show the lack of facets and the presence of fatigue striations (arrows). The approximate
∆K = ~2.7 MPa√m for all figures.
a
rn

4.4 Paris Law Fatigue Crack Growth:


u

Based on fatigue crack growth curves for the range of da/dN between ~10-6 and 10-9 m/cycle
Jo

(Figure 10), a bilinear log-log relation between da/dN and ∆K can be observed. Such
behavior is commonly observed for basketweave or lamellar microstructure (α+β) titanium
alloys [23-28, 52, 62], whereby the bilinear transition occurs when the cyclic plastic zone
size, 𝑟yc :

𝑟yc = ∆𝐾 2 ⁄12𝜋𝜎YS 2 (6)


matches the average packet/colony size [25]. However, for the CP-Ti in this study similar
packet/colony structures did not exist. Robinson et al. [34] proposed instead that the
microstructure sensitive transition should occur for CP-Ti when the cyclic plastic zone size is
on the same order of magnitude as the average grain size, d. They found the ratio of 𝑟yc /𝑑 for
the transition point ranged from less than 1 to 3 depending on grain size, with the ratio
34
increasing for decreasing grain size. While the grain sizes in the present work (~3 – 36 µm)
were much finer than in Ref. [34] (~20 – 230 µm), the trend is found to be similar with the
coarser grained materials showing 𝑟yc /𝑑 close to unity, and the finer grain materials having
𝑟yc /𝑑 closer to 3-4. (Table 7).

Table 7: Relationship between the transition stress intensity range and the ratio of the
reversed plastic zone size to grain size.

The ratio of
∆KT reversed plastic
Sample
(MPa√m) zone size to grain
size
𝒓𝐜𝐲 /𝒅

of
ASB X-Z 11.7 4.2
Z-X 11.5 2.7
HIP 730 X-Z 14.7 4.7

ro
Z-X 14.6 3.1
HIP 950 X-Z 15.1 0.7

Wrought
Z-X
T-L
L-T
16.3
10.6
11.5
-p
0.6
1.3
0.7
re
lP

4.5. Fatigue performance of CP-Ti produced by LPBF:

While there have been a handful studies examining the fatigue crack growth behavior of Ti-
a

6Al-4V manufactured by LPBF [42, 52, 89-91], this first study on CP-Ti highlights some
rn

similarities but also some distinct characteristics between the alloys. Both materials retain
significant amounts of the as-built texture after heat treatment, yet this had a weak overall
u

effect on the fatigue crack growth behavior compared to the microstructure morphology
Jo

changes [52]. Also, the increase in fatigue crack growth resistance in the microstructure
sensitive regime with increasing heat treatment temperature and α-phase grain size for both
alloys can be well explained by a blocked slip band model [52]. However, while super-β heat
treatment and the associated α-phase coarsening of Ti-6Al-4V brought about significant
amounts of crack closure effects [52], the effect of crack closure was relatively small and
consistent across all LPBF CP-Ti samples and the wrought material. Also, while near
threshold crack growth in LPBF Ti-6Al-4V has been reported to occur by a quasi-cleavage
mechanism [42], for the present LPBF CP-Ti, a striation mechanism is observed even at low
growth rates and well into the microstructure sensitive regime. While faceted crack growth
has been reported for wrought CP-Ti [35], as seen for the L-T orientation in the present study

35
(Figure 13), it involves separation of the (0002) planes that was not achieved to a large extent
for the two orthogonal build orientations examined in this study.

However, the most important distinction for LPBF CP-Ti is its sensitivity to interstitial
impurities picked up during the LPBF process, in particular nitrogen. In the present work, the
LPBF induced increase in nitrogen essentially transformed grade 2 starting powder into grade
3 final parts (Table 3). As a result, the LPBF samples had higher strength and fatigue crack
growth resistance compared to wrought grade 2 material. Furthermore, the amount of
nitrogen pickup and property change depended on the build orientation and the associated
thermal history. Overall, while the present study has demonstrated that the LPBF process can
give excellent fatigue crack growth resistance with less anisotropy than wrought material, one

of
must also be careful to monitor interstitial uptake if a specific CP-Ti grade is required for the
final part.

ro
5. Conclusions:
-p
Based on the examination of tensile and fatigue crack growth properties of laser powder bed
fusion (LPBF) fabricated CP-Ti, the following conclusions may be drawn:
re
 Hot isostatic pressing (HIP) below the β-transus temperature closed porosity while
lP

slightly coarsening the alpha grain structure, while heat treatment above the β-transus
produced a much coarser, blocky α phase grain structure. Neither HIP treatment
a

completely destroyed the texture from the LPBF process.


rn

 The presence of higher oxygen and nitrogen content gave generally higher tensile
strength and fatigue crack growth resistance for the LPBF produced samples relative
u

to wrought material. Additionally, differences in strength with respect to LPBF build


Jo

orientation were caused by differences in the nitrogen content that were attributed to
different thermal histories during fabrication.
 Increasing the α grain size by HIP treatments gave lower tensile strengths, higher
fatigue thresholds, and better fatigue crack growth resistance at growth rates less than
~ 10-7 m/cycle.
 At higher growth rates, the fatigue crack growth resistance became microstructure
insensitive and a bilinear fatigue crack growth curve was observed in the stage-II
Paris region. The transition to microstructure insensitive fatigue crack growth
occurred when the cyclic plastic zone size was on a similar order of magnitude to the
grain size.

36
 Build orientation had a negligible impact on fatigue crack growth resistance of LPBF
produced samples, while sample orientation had a significant effect for wrought
samples. The strong anisotropy observed for the wrought material was attributed to
the strong crystallographic texture. The lack of fatigue crack growth rate anisotropy
suggests potentially superior performance for LPBF produced components in real
world applications relative to components machined from commercially available
wrought material.

CRediT Authorship Contribution Statement

M. Tarik Hasib: Conceptualisation, Methodology, Investigation, Formal Analysis,

of
Visualization, Writing - Original Draft, Review & Editing. Halsey E. Ostergaard:
Methodology, Investigation, Writing - Review & Editing. Qian Liu: Methodology,

ro
Investigation, Writing - Review & Editing. Xiaopeng Li: Conceptualisation, Methodology,
Supervision, Writing - Review & Editing. Jamie J. Kruzic: Conceptualisation,
-p
Methodology, Formal Analysis, Project administration, Supervision, Writing - Review &
Editing.
re
Declaration of interests
lP

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

☒ The authors declare the following financial interests/personal relationships which may be
a

considered as potential competing interests:


u rn

M. Tarik Hasib reports financial support was provided by Australian Government Research Training
Program Scholarship.
Jo

Acknowledgements
The authors acknowledge Microscopy Australia at the Electron Microscope Unit (EMU)
within the Mark Wainwright Analytical Centre (MWAC) at UNSW Sydney for the use of
their facilities and for scientific and technical assistance. MTH also acknowledges financial
support for this study through the Australian Government Research Training Program
Scholarship.

37
References
[1] G. Lütjering, J.C. Williams, Titanium, Springer Berlin Heidelberg 2007.
[2] C. Leyens, M. Peters, Titanium and titanium alloys: fundamentals and applications, John
Wiley & Sons 2003.
[3] A. Fukuda, M. Takemoto, T. Saito, S. Fujibayashi, M. Neo, D.K. Pattanayak, T.
Matsushita, K. Sasaki, N. Nishida, T. Kokubo, T. Nakamura, Osteoinduction of porous Ti
implants with a channel structure fabricated by selective laser melting, Acta Biomaterialia
7(5) (2011) 2327-2336.
[4] D.K. Pattanayak, A. Fukuda, T. Matsushita, M. Takemoto, S. Fujibayashi, K. Sasaki, N.
Nishida, T. Nakamura, T. Kokubo, Bioactive Ti metal analogous to human cancellous bone:
Fabrication by selective laser melting and chemical treatments, Acta Biomaterialia 7(3)
(2011) 1398-1406.
[5] B. Wysocki, P. Maj, A. Krawczyńska, K. Rożniatowski, J. Zdunek, K.J. Kurzydłowski,
W. Święszkowski, Microstructure and mechanical properties investigation of CP titanium
processed by selective laser melting (SLM), J Mater Process Tech 241 (2017) 13-23.

of
[6] C. Kusuma, S.H. Ahmed, A. Mian, R. Srinivasan, Effect of Laser Power and Scan Speed
on Melt Pool Characteristics of Commercially Pure Titanium (CP-Ti), J Mater Eng Perform

ro
26(7) (2017) 3560-3568.
[7] N. Kang, H. Yuan, P. Coddet, Z. Ren, C. Bernage, H. Liao, C. Coddet, On the texture,
phase and tensile properties of commercially pure Ti produced via selective laser melting
-p
assisted by static magnetic field, Materials Science and Engineering: C 70 (2017) 405-407.
[8] D.D. Gu, W. Meiners, K. Wissenbach, R. Poprawe, Laser additive manufacturing of
metallic components: materials, processes and mechanisms, International Materials Reviews
re
57(3) (2012) 133-164.
[9] H. Attar, M. Calin, L. Zhang, S. Scudino, J. Eckert, Manufacture by selective laser
melting and mechanical behavior of commercially pure titanium, Materials Science and
lP

Engineering: A 593 (2014) 170-177.


[10] E.C. Santos, K. Osakada, M. Shiomi, Y. Kitamura, F. Abe, Microstructure and
mechanical properties of pure titanium models fabricated by selective laser melting,
a

Proceedings of the Institution of Mechanical Engineers, Part C: Journal of Mechanical


Engineering Science 218(7) (2004) 711-719.
rn

[11] K. Yamanaka, W. Saito, M. Mori, H. Matsumoto, A. Chiba, Preparation of weak-


textured commercially pure titanium by electron beam melting, Additive Manufacturing 8
(2015) 105-109.
u

[12] D. Gu, Y.-C. Hagedorn, W. Meiners, G. Meng, R.J.S. Batista, K. Wissenbach, R.


Jo

Poprawe, Densification behavior, microstructure evolution, and wear performance of


selective laser melting processed commercially pure titanium, Acta Mater 60(9) (2012) 3849-
3860.
[13] X.P. Li, J. Van Humbeeck, J.P. Kruth, Selective laser melting of weak-textured
commercially pure titanium with high strength and ductility: A study from laser power
perspective, Mater Design 116 (2017) 352-358.
[14] H. Attar, M.J. Bermingham, S. Ehtemam-Haghighi, A. Dehghan-Manshadi, D. Kent,
M.S. Dargusch, Evaluation of the mechanical and wear properties of titanium produced by
three different additive manufacturing methods for biomedical application, Mater. Sci. Eng.
A 760 (2019) 339-345.
[15] T.W. Na, W.R. Kim, S.M. Yang, O. Kwon, J.M. Park, G.H. Kim, K.H. Jung, C.W. Lee,
H.K. Park, H.G. Kim, Effect of laser power on oxygen and nitrogen concentration of
commercially pure titanium manufactured by selective laser melting, Mater. Charact. 143
(2018) 110-117.
[16] H. Wei, J. Mazumder, T. DebRoy, Evolution of solidification texture during additive
manufacturing, Scientific reports 5 (2015) 16446.

38
[17] G. Lütjering, Influence of processing on microstructure and mechanical properties of
(α+β) titanium alloys, Materials Science and Engineering: A 243(1) (1998) 32-45.
[18] H. Galarraga, R.J. Warren, D.A. Lados, R.R. Dehoff, M.M. Kirka, P. Nandwana, Effects
of heat treatments on microstructure and properties of Ti-6Al-4V ELI alloy fabricated by
electron beam melting (EBM), Materials Science and Engineering: A 685 (2017) 417-428.
[19] S. Bahl, S. Suwas, K. Chatterjee, The importance of crystallographic texture in the use
of titanium as an orthopedic biomaterial, RSC Advances 4(72) (2014) 38078-38087.
[20] I. Bantounas, D. Dye, T.C. Lindley, The effect of grain orientation on fracture
morphology during high-cycle fatigue of Ti–6Al–4V, Acta Mater 57(12) (2009) 3584-3595.
[21] W. Evans, J. Jones, M. Whittaker, Texture effects under tension and torsion loading
conditions in titanium alloys, Int J Fatigue 27(10) (2005) 1244-1250.
[22] V. Sinha, M. Mills, J. Williams, Crystallography of fracture facets in a near-alpha
titanium alloy, Metallurgical and Materials Transactions A 37(6) (2006) 2015-2026.
[23] P.E. Irving, C.J. Beevers, Microstructural influences on fatigue crack growth in Ti-6Al-
4V, Materials Science and Engineering 14(3) (1974) 229-238.

of
[24] G. Yoder, L. Cooley, T. Crooker, Fatigue crack propagation resistance of beta-annealed
Ti-6AI-4V alloys of differing interstitial oxygen contents, Metallurgical Transactions A 9(10)
(1978) 1413-1420.

ro
[25] G.R. Yoder, L.A. Cooley, T.W. Crooker, Quantitative analysis of microstructural effects
on fatigue crack growth in widmanstätten Ti-6A1-4V and Ti-8Al-1Mo-1V, Engineering
Fracture Mechanics 11(4) (1979) 805-816.
-p
[26] G.R. Yoder, L.A. Cooley, T.W. Crooker, Observations on microstructurally sensitive
fatigue crack growth in a widmanstätten Ti-6Al-4V alloy, Metallurgical Transactions A 8(11)
re
(1977) 1737-1743.
[27] K. Ravichandran, E. Dwarakadasa, Fatigue crack growth transitions in Ti-6Al-4V alloy,
Scripta metallurgica 23(10) (1989) 1685-1690.
lP

[28] H. Wang, A study on the change of fatigue fracture mode in two titanium alloys, Fatigue
& Fracture of Engineering Materials & Structures 21(9) (1998) 1077-1087.
[29] G.J. Baxter, W.M. Rainforth, L. Grabowski, TEM observations of fatigue damage
accumulation at the surface of the near-alpha titanium alloy IMI 834, Acta Mater. 44(9)
a

(1996) 3453-3463.
rn

[30] M.R. Bache, W.J. Evans, V. Randle, R.J. Wilson, Characterization of mechanical
anisotropy in titanium alloys, Mater. Sci. Eng. A 257(1) (1998) 139-144.
[31] M. Bache, W. Evans, B. Suddell, F. Herrouin, The effects of texture in titanium alloys
u

for engineering components under fatigue, Int J Fatigue 23 (2001) 153-159.


[32] L. Li, Z. Zhang, G. Shen, The Effect of Grain Size on Fatigue Crack Propagation in
Jo

Commercial Pure Titanium Investigated by Acoustic Emission, J Mater Eng Perform 24


(2015) 2720-2729.
[33] Y. Mine, S. Ando, K. Takashima, Crystallographic fatigue crack growth in titanium
single crystals, Materials Science and Engineering: A 528(25-26) (2011) 7570-7578.
[34] J.L. Robinson, C.J. Beevers, The Effects of Load Ratio, Interstitial Content, and Grain
Size on Low-Stress Fatigue-Crack Propagation in α-Titanium, Metal Science Journal 7(1)
(1973) 153-159.
[35] C.M. Ward-Close, C.J. Beevers, The influence of grain orientation on the mode and rate
of fatigue crack growth inα-titanium, Metallurgical Transactions A 11(6) (1980) 1007-1017.
[36] Z.N. Ismarrubie, H. Yussof, M. Sugano, Fatigue damage mechanism of titanium in
vacuum and in air, in: A.K. Makhtar, H. Yussof, H. AlAssadi, L.C. Yee (Eds.), International
Symposium on Robotics and Intelligent Sensors 2012, Elsevier Science Bv, Amsterdam,
2012, pp. 1559-1565.
[37] S. Ando, T. Sakamoto, Y. Ikejiri, M. Tsushida, H. Tonda, Crack orientation dependence
of fatigue behavior in titanium single crystals by thin sheet plain bending, in: S.B. Lee, Y.J.

39
Kim (Eds.), Experimental Mechanics in Nano and Biotechnology, Pts 1 and 2, Trans Tech
Publications Ltd, Stafa-Zurich, 2006, pp. 967-970.
[38] K. Tokaji, T. Ogawa, K. Ohya, The effect of grain size on small fatigue crack growth in
pure titanium, International Journal of Fatigue 16(8) (1994) 571-578.
[39] N.G. Turner, W.T. Roberts, Fatigue behavior of titanium, Transaction of the
Metallurgical Society of AIME 242 (1968) 1223-1230.
[40] S.H. Ai, Z.G. Wang, Y.B. Xia, The fatigue deformation and fracture characteristics of
coarse grained polycrystalline α-titanium, Scripta Metall. 19(9) (1985) 1089-1093.
[41] M. Sugano, C.M. Gilmore, A crystallographic study of fatigue damage in titanium,
Metallurgical Transactions A 11(4) (1980) 559-563.
[42] T.H. Becker, N.M. Dhansay, G.M.T. Haar, K. Vanmeensel, Near-threshold fatigue crack
growth rates of laser powder bed fusion produced Ti-6Al-4V, Acta Mater. 197 (2020) 269-
282.
[43] Y. Okazaki, A. Ishino, Microstructures and Mechanical Properties of Laser-Sintered
Commercially Pure Ti and Ti-6Al-4V Alloy for Dental Applications, Materials 13(3) (2020)

of
609.
[44] J.-P. Kruth, P. Mercelis, J.V. Vaerenbergh, L. Froyen, M. Rombouts, Binding
mechanisms in selective laser sintering and selective laser melting, Rapid Prototyping J 11(1)

ro
(2005) 26-36.
[45] L. Zhang, T. Sercombe, Selective laser melting of low-modulus biomedical Ti-24Nb-
4Zr-8Sn alloy: effect of laser point distance, Key Engineering Materials, Trans Tech Publ,
2012, pp. 226-233. -p
[46] A.B. Spierings, M. Schneider, R. Eggenberger, Comparison of density measurement
re
techniques for additive manufactured metallic parts, Rapid Prototyping J 17(5) (2011) 380-
386.
[47] ASTM Standard E8/E8M-16ae1 Standard Test Methods for Tension Testing of Metallic
lP

Materials, ASTM International, West Conshohocken, PA; ASTM International, 2016. DOI:
10.1520/E0008_E0008M-16AE01, www.astm.org.
[48] ASTM Standard E647-15e1, Standard Test Method for Measurement of Fatigue Crack
Growth Rates, ASTM International, West Conshohocken, Pennsylvania, USA, 2015, DOI:
a

10.1520/E0647-1515E1501, www.astm.org.
rn

[49] ISO/ASTM52921-13(2019) Standard Terminology for Additive


Manufacturing;Coordinate Systems and Test Methodologies, ASTM International, West
Conshohocken, PA, 2019. doi: https://doi.org/10.1520/ISOASTM52921-13R19.
u

[50] J.E. Hilliard, Conversion of intercept density to grain size, Met. Prog 85(5) (1964) 99.
[51] ISO standard ISO/DIS 21920, Geometrical Product Specifications (GPS)--surface
Jo

Texture: Profile Method--rules and Procedures for the Assessment of Surface Texture
Switzerland,DOI: 10.31030/3137774, www.iso.org, Geometrical Product Specifications
(GPS)--surface Texture: Profile Method--rules and Procedures for the Assessment of Surface
Texture Switzerland,DOI: 10.31030/3137774, www.iso.org, 2020(E).
[52] M. Tarik Hasib, H.E. Ostergaard, X. Li, J.J. Kruzic, Fatigue crack growth behavior of
laser powder bed fusion additive manufactured Ti-6Al-4V: Roles of post heat treatment and
build orientation, Int J Fatigue (2020) 105955.
[53] M. Simonelli, Y.Y. Tse, C. Tuck, On the Texture Formation of Selective Laser Melted
Ti-6Al-4V, Metallurgical and Materials Transactions A 45(6) (2014) 2863-2872.
[54] N. Gey, M. Humbert, Characterization of the variant selection occurring during the
α→β→α phase transformations of a cold rolled titanium sheet, Acta Mater 50(2) (2002) 277-
287.
[55] S. Nourbakhsh, T.D. O'Brien, Texture formation and transition in Cold-rolled titanium,
Materials Science and Engineering 100 (1988) 109-114.

40
[56] Z.S. Zhu, J.L. Gu, R.Y. Liu, N.P. Chen, M.G. Yan, Variant selection and its effect on
phase transformation textures in cold rolled titanium sheet, Materials Science and
Engineering: A 280(1) (2000) 199-203.
[57] C. Qiu, N.J.E. Adkins, M.M. Attallah, Microstructure and tensile properties of
selectively laser-melted and of HIPed laser-melted Ti–6Al–4V, Materials Science and
Engineering: A 578 (2013) 230-239.
[58] J. Yang, H. Yu, J. Yin, M. Gao, Z. Wang, X. Zeng, Formation and control of martensite
in Ti-6Al-4V alloy produced by selective laser melting, Materials & Design 108 (2016) 308-
318.
[59] G.M. Ter Haar, T.H. Becker, Selective Laser Melting Produced Ti-6Al-4V: Post-Process
Heat Treatments to Achieve Superior Tensile Properties, Materials 11(1) (2018) 146.
[60] S.-S. Rui, L.-S. Niu, H.-J. Shi, S. Wei, C.C. Tasan, Diffraction-based misorientation
mapping: A continuum mechanics description, Journal of the Mechanics and Physics of
Solids 133 (2019) 103709.
[61] U. Bathini, T.S. Srivatsan, A. Patnaik, T. Quick, A Study of the Tensile Deformation and

of
Fracture Behavior of Commercially Pure Titanium and Titanium Alloy: Influence of
Orientation and Microstructure, J Mater Eng Perform 19(8) (2010) 1172-1182.
[62] R.J.H. Wanhill, R. Galatolo, C. Looije, Fractographic and microstructural analysis of

ro
fatigue crack growth in a Ti-6Al-4V fan disc forging, Int J Fatigue 11(6) (1989) 407-416.
[63] N. Bozzolo, N. Dewobroto, T. Grosdidier, F. Wagner, Texture evolution during grain
growth in recrystallized commercially pure titanium, Materials Science and Engineering: A
397(1) (2005) 346-355. -p
[64] S.S. da Rocha, G.L. Adabo, L.G. Vaz, G.E.P. Henriques, Effect of thermal treatments on
re
tensile strength of commercially cast pure titanium and Ti-6Al-4V alloys, Journal of
Materials Science: Materials in Medicine 16(8) (2005) 759-766.
[65] H. Nasiri-Abarbekoh, A. Ekrami, A.A. Ziaei-Moayyed, Impact of phase transformation
lP

on mechanical properties anisotropy of commercially pure titanium, Mater Design


37(Supplement C) (2012) 223-227.
[66] W.L. Finlay, J.A. Snyder, Effects of three interstitial solutes (nitrogen, oxygen, and
carbon) on the mechanical properties of high-purity, alpha titanium, JOM 2(2) (1950) 277-
a

286.
rn

[67] A. Issariyapat, P. Visuttipitukul, T. Song, A. Bahador, J. Umeda, M. Qian, K. Kondoh,


Tensile properties improvement by homogenized nitrogen solid solution strengthening of
commercially pure titanium through powder metallurgy process, Mater Charact 170 (2020)
u

110700.
[68] K. Kondoh, A. Issariyapat, J. Umeda, P. Visuttipitukul, Selective laser-melted titanium
Jo

materials with nitrogen solid solutions for balanced strength and ductility, Materials Science
and Engineering: A 790 (2020) 139641.
[69] K. Kondoh, B. Sun, S. Li, H. Mai, J. Umeda, Experimental and theoretical analysis of
nitrogen solid-solution strengthening of PM titanium, International Journal of Powder
Metallurgy 50 (2014) 35-40.
[70] P. Kwasniak, H. Garbacz, K. Kurzydlowski, Solid solution strengthening of hexagonal
titanium alloys: Restoring forces and stacking faults calculated from first principles, Acta
Mater 102 (2016) 304-314.
[71] B. Sun, S. Li, H. Imai, T. Mimoto, J. Umeda, K. Kondoh, Fabrication of high-strength Ti
materials by in-process solid solution strengthening of oxygen via P/M methods, Materials
Science and Engineering: A 563 (2013) 95-100.
[72] Q. Yu, L. Qi, T. Tsuru, R. Traylor, D. Rugg, J. Morris, M. Asta, D. Chrzan, A.M. Minor,
Origin of dramatic oxygen solute strengthening effect in titanium, Science 347(6222) (2015)
635-639.

41
[73] T. Ando, K. Nakashima, T. Tsuchiyama, S. Takaki, Microstructure and mechanical
properties of a high nitrogen titanium alloy, Materials Science and Engineering: A 486(1)
(2008) 228-234.
[74] T.-W. Na, W.R. Kim, S.-M. Yang, O. Kwon, J.M. Park, G.-H. Kim, K.-H. Jung, C.-W.
Lee, H.-K. Park, H.G. Kim, Effect of laser power on oxygen and nitrogen concentration of
commercially pure titanium manufactured by selective laser melting, Mater Charact 143
(2018) 110-117.
[75] R. Labusch, A Statistical Theory of Solid Solution Hardening, physica status solidi (b)
41(2) (1970) 659-669.
[76] S. Kariya, M. Fukuo, J. Umeda, K. Kondoh, Quantitative Analysis on Light Elements
Solution Strengthening in Pure Titanium Sintered Materials by Labusch Model Using
Experimental Data, Mater Trans 60(2) (2019) 263-268.
[77] K. Kondoh, E. Ichikawa, A. Issariyapat, K. Shitara, J. Umeda, B. Chen, S. Li, Tensile
property enhancement by oxygen solutes in selectively laser melted titanium materials
fabricated from pre-mixed pure Ti and TiO2 powder, Materials Science and Engineering: A

of
795 (2020) 139983.
[78] Moreno-Valle, W. Pachla, M. Kulczyk, I. Sabirov, A. Hohenwarter, Anisotropy of
tensile and fracture behavior of pure titanium after hydrostatic extrusion, Mater Trans 60(10)

ro
(2019) 2160-2167.
[79] S. Suresh, R.O. Ritchie, A geometric model for fatigue crack closure induced by fracture
surface roughness, Metallurgical Transactions A 13(9) (1982) 1627-1631.
-p
[80] R. Ritchie, S. Suresh, Some considerations on fatigue crack closure at near-threshold
stress intensities due to fracture surface morphology, Metallurgical Transactions A 13(5)
re
(1982) 937-940.
[81] S. Suresh, Fatigue crack deflection and fracture surface contact: Micromechanical
models, Metallurgical Transactions A 16(1) (1985) 249-260.
lP

[82] T. Ogawa, K. Tokaji, K. Ohya, THE EFFECT OF MICROSTRUCTURE AND


FRACTURE SURFACE ROUGHNESS ON FATIGUE CRACK PROPAGATION IN A Ti-
6A1-4V ALLOY, Fatigue & Fracture of Engineering Materials & Structures 16(9) (1993)
973-982.
a

[83] K. Sadananda, A.K. Vasudevan, Fatigue crack growth behavior of titanium alloys, Int J
rn

Fatigue 27(10) (2005) 1255-1266.


[84] S. Kikuchi, T. Mori, H. Kubozono, Y. Nakai, M.O. Kawabata, K. Ameyama, Evaluation
of near-threshold fatigue crack propagation in harmonic-structured CP titanium with a
u

bimodal grain size distribution, Engineering Fracture Mechanics 181(Supplement C) (2017)


77-86.
Jo

[85] X.-D. Li, Dislocation pile-up model of fatigue thresholds for 2024- and 7075-alike
aluminium alloys, Theoretical and Applied Fracture Mechanics 24(2) (1996) 165-179.
[86] X.-D. Li, L. Edwards, Theoretical modelling of fatigue threshold for aluminium alloys,
Eng. Fract. Mech. 54(1) (1996) 35-48.
[87] L. Lawson, E.Y. Chen, M. Meshii, Near-threshold fatigue: a review, International
Journal of Fatigue 21 (1999) S15-S34.
[88] K. Tanaka, Y. Nakai, M. Yamashita, Fatigue growth threshold of small cracks, Int. J.
Fract. 17(5) (1981) 519-533.
[89] P. Edwards, M. Ramulu, Effect of build direction on the fracture toughness and fatigue
crack growth in selective laser melted Ti‐ 6Al‐ 4 V, Fatigue & Fracture of Engineering
Materials & Structures 38(10) (2015) 1228-1236.
[90] S. Leuders, M. Thöne, A. Riemer, T. Niendorf, T. Tröster, H.A. Richard, H.J. Maier, On
the mechanical behaviour of titanium alloy TiAl6V4 manufactured by selective laser melting:
Fatigue resistance and crack growth performance, Int J Fatigue 48 (2013) 300-307.

42
[91] P. Kumar, U. Ramamurty, Microstructural optimization through heat treatment for
enhancing the fracture toughness and fatigue crack growth resistance of selective laser melted
Ti6Al4V alloy, Acta Mater 169 (2019) 45-59.

of
ro
-p
re
a lP
u rn
Jo

43
Graphical Abstract

of
ro
-p
re
alP
u rn
Jo

44
Highlights

 Fatigue crack growth rates measured for laser powder bed fusion CP-Ti
 Interstitial pickup during fabrication raised tensile strength and fatigue resistance
 Microstructure sensitive growth when cyclic plastic zone - grain size ratio is ≲1-4
 α'/α lath thickness and interstitial content controlled fatigue crack growth rates
 Build orientation had negligible effect on fatigue crack growth

of
ro
-p
re
a lP
u rn
Jo

45

You might also like