You are on page 1of 16

International Journal of Fatigue 127 (2019) 58–73

Contents lists available at ScienceDirect

International Journal of Fatigue


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

Low cycle fatigue properties of Ti-6Al-4V alloy fabricated by high-power T


laser directed energy deposition: Experimental and prediction

Y.M. Rena,b,c, X. Lina,b, , P.F. Guoa,b, H.O. Yanga,b, H. Tana,b, J. Chena,b, J. Lia, Y.Y. Zhanga,b,

W.D. Huanga,b,
a
State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an, Shaanxi 710072, PR China
b
Key Laboratory of Metal High Performance Additive Manufacturing and Innovative Design, MIIT China, Northwestern Polytechnical University, Xi’an, Shaanxi 710072,
PR China
c
The Shaanxi Key Laboratory of Photoelectric Functional Materials and Devices, Xi’an Technological University, Xi’an 710021, PR China

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

Keywords: The low cycle fatigue (LCF) properties of Ti-6Al-4V parts fabricated by high-power laser directed energy de-
Low cycle fatigue position were investigated. The results show that the parts after solution treatment and aging exhibit superior
Titanium alloy LCF lives to those of other reported AM Ti-6Al-4V parts, as well as are comparable to those of the wrought
Laser directed energy deposition counterparts at intermediate strain amplitudes (from 0.8% to 1.1%). Cyclic softening behaviors were found at
Additive manufacturing
various strain amplitudes (from 0.65% to 1.7%). In addition, a microstructure-based multistage fatigue model
Fatigue model
was used to predict their LCF lives and shows good agreement with the experimental data.

1. Introduction the LCF properties and fracture behavior of AM Ti-6Al-4V parts are still
not fully understood.
Laser directed energy deposition additive manufacturing (AM) is a Laser solid forming (LSF), a laser directed energy deposition tech-
near-net shape production technology can directly fabricate large and nology based on simultaneous powder feeding, which is also known as
complex geometry titanium alloys parts used in aerospace and auto- laser metal deposition (LMD) or Laser Engineered Net Shaping (LENS)
motive fields [1,2]. Titanium alloys, particularly Ti-6Al-4V, are in- [1,24,25], has advantages in the fabrication of large and complex parts.
creasingly being used for structural applications in aerospace because of Recently, with the development of the energy output stability for lasers,
their high specific strength and excellent corrosion resistance [3]. In high-power (≥4000 W) lasers are used for LSF technology to improve
general, the tensile properties of AM Ti-6Al-4V parts [4–7] are com- deposition efficiency, and much attention has been paid in the manu-
parable to those of their wrought counterparts. However, the low cycle facturing of large-scale complex titanium alloy parts [26,27]. Mean-
fatigue (LCF) properties of AM Ti-6Al-4V parts [8–10] are inferior to the while, the microstructure of the high-power LSF parts is different from
wrought standard, which restricts its widely used in industrial appli- that of the conventional low-power conditions [28,29], thus it will re-
cation. To improve the working reliability and trustworthiness of AM sult in different low-cycle fatigue property. Generally, AM parts fabri-
Ti-6Al-4V parts, their LCF properties should be improved. cated by high power lasers typically have a large scale rough surface
Several authors [11,12] have reported the LCF properties of AM Ti- due to large laser beam and large layer thickness [1]. In addition, the
6Al-4V parts and they believed that the microstructure, defects (e.g., porosity of deposited parts depends on the combination of deposition
pores, lack-of-fusion, and unmelted powders), residual stress, and rough parameters (e.g., laser power, scanning velocity, layer thickness) [1].
surface can strongly affect their LCF properties. For example, the lack- Importantly, high-power AM processes often induce coarse prior-β
of-fusion defects decrease the LCF property more severely than the grains with thick α-laths in AM parts [30,31], which would strongly
pores and unmelted powders [11]. Thus, optimizing AM processing affect LCF and tensile properties. In this work, the LCF properties of Ti-
parameters [13,14] and applying hot isostatic pressing [15–17] would 6Al-4V parts fabricated by high power (7600 W) LSF were studied. The
strongly enhance the fatigue properties of AM parts. Although many cyclic deformation behaviors of LSF Ti-6Al-4V specimens were in-
investigators [1,18–23] reviewed the fatigue behavior of AM Ti-6Al-4V vestigated. In addition, a microstructure-based multistage fatigue
parts, their primary concern is high cycle fatigue properties. However, model was used to predict the low cycle fatigue life of LSF Ti-6Al-4V


Corresponding authors at: State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an, Shaanxi 710072, PR China.
E-mail addresses: xlin@nwpu.edu.cn (X. Lin), huang@nwpu.edu.cn (W.D. Huang).

https://doi.org/10.1016/j.ijfatigue.2019.05.035
Received 8 March 2019; Received in revised form 21 May 2019; Accepted 30 May 2019
Available online 31 May 2019
0142-1123/ © 2019 Elsevier Ltd. All rights reserved.
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

alloy. roughness. That is, after polishing, all remaining grinding and polishing
marks should be longitudinal. Before the LCF tests, tensile tests were
2. Experimental procedures also conducted at room temperature using an INSTRON 3382 universal
testing system at a strain rate of 1.7 × 10-3 s−1. We conducted the
2.1. Deposited parts preparation tensile tests three times under a given condition (V900 and V1500,
respectively). The tensile test samples were cut from the deposited
Deposited parts (140 × 70 × 90 mm3) of Ti-6Al-4V alloy were fab- samples (140 × 70 × 90 mm3) with gauge diameter of 4 mm, gauge
ricated via LSF [25], using an LSF-V system (Rofin DC80 8 kW CO2) length of 15 mm and total length of 73 mm.
under an argon atmosphere (oxygen content less than 50 ppm). The Ti-
6Al-4V powder was produced by plasma rotating electrode process with 2.3. LCF model and analysis
a spherical particle size between 45 and 325 μm. The average chemical
composition of the Ti-6Al-4V powder (wt%) was measured as Al-6.15, Generally, there are many methods for evaluating low cycle fatigue
V-4.05, O-0.15, H-0.004, N-0.01, C-0.005, Fe-0.09, Si-0.02 and Ti bal- life and the two well-known methods were widely used. One is the
ance. The laser processing parameters are as follows: laser power universal slope method, and the other is the local strain approach.
7.6 kW, beam diameter 6 mm, powder feed rate 500 g/h, overlap 40%, Manson [33] first proposed the universal slope method to evaluate the
Z-axis increment 0.8 mm, and scanning velocities 900 and 1500 mm/ low cycle fatigue data or strain-life curve. In addition, the slopes of the
min. To reduce the anisotropy of the mechanical properties, a cross- plastic and elastic lines are universalized for all materials (c = −0.6;
directional (N-layer, 45°; and (N + 1) layer, 135°) laser scanning b = −0.12), while σf′and εf′are dependent on tensile strength and true
strategy was used. To improve the ductility and uniformity of LSF fracture strength [33,34]. Later, to improve the prediction accuracy for
samples, the solution treatment and aging (STA) were conducted at LCF life, a modified universal slope method (c = −0.56, b = −0.09)
950 °C for 2 h followed by air cooling (AC) and 550 °C for 4 h followed was developed by Muralidharan and Manson [35]. A more detailed
by AC. After post-LSF heat treatment, we remove the surface oxygen description of estimation method for strain-life curves can be found in
layer and machined the samples far away from the surface oxygen layer M. Ricotta’s work [34].
of the deposits, and the fatigue samples were cut by electrical discharge The local strain approach relates the applied strain amplitude to the
machining (EDM) in the deposited parts. The oxygen content of as- number of reversals to failure, according to the Basquin, Manson and
deposited parts after heat treatment are 0.18 wt%, indicating the Coffin equations [36–38]. The classical strain-controlled LCF approach
oxygen content slightly increased in the deposited parts after STA was used to predict the fatigue life for the initiation and early growth of
process. The residual stress of the heat-treated sample is smaller than a crack within the strain field associated with the fully plastic region
that of the as-deposited parts, this result supported by the study from ahead of a stress concentration [39]. In general, the total strain am-
Leuders et al. [32]. In addition, the residual stress for the fatigue spe- plitude, Δε/2, separated elastic and plastic strain amplitude (Eq. (1)).
cimen will change during this EDM process. Therefore, in this work, we To evaluate the elastic strain component of fatigue lives, the Basquin
did not consider the effect of residual stress on fatigue lives for STA equation (Eq. (2)) was used. LCF behavior was analyzed using the
specimens. A more detailed description of the LSF Ti-6Al-4V experi- Coffin-Manson relation (Eq. (3)) between the plastic strain amplitude
mental and deposition patterns can be found in Ref. [25]. Δεp/2 and the number of reversals to failure (2Nf). According to Eqs.
(1)–(3), one can obtain the total relation (Eq. (4)) between the strain
2.2. Low cycle fatigue tests amplitude and fatigue life. The cyclic stress-plastic strain amplitude
curve is characterized by the relationship presented in Eq. (5).
Strain-controlled LCF tests were carried out at room temperature in
Δε /2 = Δεe/2 + Δεp/2 (1)
fully reversed mode (strain ratio Rε = −1, f = 0.3 Hz, triangular wa-
veform) at constant total strain amplitudes (Δε/2 = 0.55%, 0.65%, σf′
0.7%, 0.8%, 1.1%, 1.4% and 1.7%, respectively) using an INSTRON Δεe/2 = (2Nf )b
E (2)
8802 with extensometer of 12.5 mm gauge length. LCF samples with
gauge diameter of 6 mm, gauge length of 15 mm and total length of Δεp/2 = εf′ (2Nf )c (3)
73 mm were extracted from the LSF Ti-6Al-4V deposited parts by EDM
σf′
in accordance with Chinese standard for low cycle fatigue tests GB/T Δε /2 = (2Nf )b + εf′ (2Nf )c
E (4)
15248-2008. The gauge section of each fatigue specimen was polished
along the longitudinal axis to reduce surface roughness effects on fa- Δσ /2 = K ′ (Δεp/2)n′ (5)
tigue behavior. We did not consider the surface roughness in this work,
because the fatigue specimen was machined by EDM from the heat- where Δε/2 is the total strain amplitude, Δσ/2 is the stress amplitude,
treated deposited parts. In addition, the specimen have strictly re- Δεe/2 is the elastic strain amplitude, Δεp/2 is the plastic strain ampli-
quirement for surface roughness (0.32 μm) for the fatigue specimen. tude, σf′ is the fatigue strength coefficient, εf′ is the fatigue ductility
Twenty fatigue specimens for V900 and 21 for V1500 were con- coefficient, b is the fatigue strength exponent, c is the fatigue ductility
ducted at seven strain amplitudes. At the beginning of the LCF tests, we exponent, K ′ is the cyclic strength coefficient, n′ is the cyclic strain
conducted two V900 specimens for evaluating LCF lives at a strain hardening exponent, E is the Young's modulus (∼117 GPa for LSF Ti-
amplitude of 0.55% and found that the specimens were subjected to 6Al-4V alloy), and 2Nf is the number of reversals to failure.
little plastic strain from their hysteresis loops, showing an approxi-
mately linear relation. Thus, we increased the strain amplitudes (from 3. Results
0.65%) for the following LCF tests for both V900 and V1500 specimens.
For a given strain amplitude, a minimum of three specimens was tested 3.1. Low cycle fatigue properties
to increase the fatigue data reliability, except for two specimens fati-
gued at high strain amplitude of 1.7% owing to their highly unstable Fig. 1 shows the low cycle fatigue properties of the STA LSF Ti-6Al-
fatigue test processes. LCF lives of LSF Ti-6Al-4V specimens were 4V samples at two scanning velocities (900 and 1500 mm/min). By
counted after the specimen fractured. Both the loading stress axis of the comparing the LCF properties for V900 and V1500 samples, we found
LCF and tensile samples are parallel to the deposition direction (Z-axis) that the LCF properties at scanning velocities 900 and 1500 mm/min
of the LSF components. Note that all fatigue specimens were polished are not easy to distinguish (Fig. 1a). This small difference of LCF
along the longitudinal direction to impart a maximum of 0.4 μm surface properties existing in V900 and V1500 samples was due to the scatter in

59
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 1. Low cycle fatigue properties of the STA LSF Ti-6Al-4V samples at two scanning velocities. (a) Sample position in deposited parts (140 × 70 × 90 mm3) and
the geometry of LCF samples, (b) LCF data at scanning velocities of 900 and 1500 mm/min, (c,d) Plastic amplitude range (Δεp/2) and reversals to failure (2Nf) at
scanning velocities of 900 and 1500 mm/min, (e,f) Stress amplitude and plastic amplitude at scanning velocities of 900 and 1500 mm/min. The forging equation is
derived from Ref. [40].

Table 1 STA LSF Ti-6Al-4V samples are comparable to the values of wrought
Strain-controlled fatigue parameters of the STA LSF Ti-6Al-4V samples. standard [40], whereas at strain amplitudes (Δε/2) between 1.1% and
σf′ (MPa) b εf′ c k′ (MPa) n’ 1.7% the LCF properties are smaller than those of wrought standard.
The Z-direction of deposited parts and the loading direction for the
V900 1134.90 −0.05217 0.20621 −0.57527 984.12 0.04781 fatigue specimen are shown in Fig. 1b. According to the Coffin-Manson
V1500 1106.82 −0.04474 0.21597 −0.60018 1033.20 0.04868
and Basquin equations, the relationships of Δε/2 − 2Nf, Δεp/2 − 2Nf,
Δεe/2 − 2Nf and Δσ/2 − Δεp/2 are depicted (see Fig. 1c–f) at scanning
Note: σf′: Fatigue strength coefficient; b: Fatigue strength exponent; εf′: Fatigue
ductility coefficient; c: Fatigue ductility exponent; k′: Cyclic strength coeffi- velocities of 900 and 1500 mm/min. From these fitting lines (see
cient; n′: Cyclic strain hardening exponent. Fig. 1c–f) of the important fatigue parameters, we can compare the
difference between LCF data in V900 and V1500 samples. The curve for
LCF data from the heterogeneous microstructure and the presence of wrought standard [40] is not a minimum bound. This curve fitted by
defects (such as pores, lack-of-fusion) in the specimens. In addition, at Coffin-Manson equation presents best fitting effect for all fatigue data.
strain amplitudes (Δε/2) between 0.8% and 1.1%, the LCF properties of The wrought standard Ti-6Al-4V alloy [40] exhibits equiaxed

60
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 2. Fatigue failure process of V1500 Ti-6Al-4V at a strain amplitude of 1.4%: (a) stress amplitude and Nf. (b) plastic strain amplitude and Nf; (c) stress and total
cycle times; (d) the magnified image at the beginning stage (at the beginning of the 40 s) corresponding to (c) marked dashed rectangle. (e) elastic modulus and Nf, (f)
the typical low cycle fatigued morphology of the V1500 sample at a strain amplitude of 1.4%.

microstructure (i.e. equiaxed α phase and β phases). For LSF Ti-6Al-4V (Fig. 2c). Note that the stress values gradually reduced during the cyclic
alloy, the microstructure has columnar prior-β grains and α-laths loading. Furthermore, this result also demonstrates that LSF Ti-6Al-4V
within the grains. Compared with LSF Ti-6Al-4V alloy, the wrought Ti- alloy displayed tension-compression asymmetry behavior during LCF
6Al-4V alloy has relatively finer grains than the LAM Ti-6Al-4V alloy. loading. For example, under the strain amplitude of 1.4% condition, the
The values of the important fatigue parameters are listed in Table 1. maximum tension stress is 918 MPa, and the minimum compression-
From Table 1, we found fatigue strength exponent (b) and fatigue stress is −993 MPa (Fig. 2d), which caused the ratio of tension to
ductility exponent (c) are close to the values (c = −0.56, b = −0.09) compression stress to not be equal to their strain ratio −1 (the real ratio
from a modified universal slope method was developed by Mur- is −0.92). The loading and unloading moduli decreased with increasing
alidharan and Manson [35]. This result demonstrates that the local Nf during the fatigue process (Fig. 2e). In addition, Fig. 2f shows a ty-
strain method is consistent with the modified universal slope method to pical macroscopic fatigued fracture of the LSF Ti-6Al-4V alloy. The
evaluate the low cycle fatigue life of LSF Ti-6Al-4V alloy. fracture surface typically exhibits inclined characters to the loading
To understand the LCF mechanism, the detailed fatigue failure axis.
process of STA LSF Ti-6Al-4V under the strain amplitude of 1.4% con-
dition was investigated (Fig. 2). The STA LSF Ti-6Al-4V shows initial 3.2. Cyclic softening behavior
cyclic softening followed by steady state stage (Fig. 2a,b) and a typical
fatigued fracture surface (inset in Fig. 2b). The trend of the stress-time Generally, the engineering alloys subjected to fatigue loading ex-
relationship is similar to the stress-Nf curves at the whole stages hibit cyclic softening or cyclic hardening phenomena [39]. A stress-Nf

61
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 3. Cyclic softening behavior of the STA LSF Ti-6Al-4V samples under different strain amplitudes: (a) Δε/2 = 0.65%, (b) Δε/2 = 0.7%, (c) Δε/2 = 0.8%, (d) Δε/
2 = 1.1%, (e) Δε/2 = 1.4% , (f) Δε/2 = 1.7%,

curve (Fig. 2a) shows that the STA LSF Ti-6Al-4V samples subjected to follows [42]:
strain-controlled LCF loads exhibit the cyclic softening phenomenon. To
investigate the cyclic softening behavior of the STA LSF Ti-6Al-4V
CSR = (Δσ max − Δσ half )/Δσ max (6)
samples, the hysteresis loop at different strain amplitudes (Fig. 3) was where CSR is the cyclic softening ratio, Δσ max
is the maximum stress
used. Fig. 3 shows the cyclic softening behavior of the STA LSF Ti-6Al- amplitude and Δσhalf is the stress amplitude at the half lifetime.
4V samples at six strain amplitudes (0.65%, 0.7%, 0.8%, 1.1%, 1.4% Fig. 4 shows that the cyclic softening ratio of the LSF Ti-6Al-4V alloy
and 1.7%). The hysteresis loops enlarged with the strain amplitudes increased dramatically (from ∼0 to 0.12) at low strain amplitudes
increased from 0.65% to 1.7%. In addition, with increasing the number (from 0.65% to 0.8%), whereas the SR increased slowly from ∼0.12 to
of cycles (Nf), the stress decreased at a given strain amplitude ∼0.17 at high strain amplitudes (from 1.1% to 1.7%). With an increase
(Fig. 3a–f). These results demonstrate that the STA LSF Ti-6Al-4V alloy in strain amplitude, the cyclic softening behavior of STA LSF Ti-6Al-4V
exhibited the cyclic softening behavior under the strain–controlled enhanced. Note that the curves of cyclic softening ratio and strain
cyclic loading conditions. In general, the initial state of material con- amplitude almost overlap in V900 and V1500 samples during the LCF
trolled the cyclic softening or hardening behavior [41]. When the initial process. These results indicate that strain amplitudes during LCF tests
material belongs to soft state (UTS/YS ≥ 1.4), cyclic hardening occurs; strongly affect the cyclic softening ratio during LCF tests. However,
however, in the hard state (UTS/YS < 1.2), cyclic softening occurs. We these results also demonstrate that the microstructures have little effect
observed cyclic softening in this study, where the ratio of UTS/ on the cyclic softening behavior. Fig. 5a and b shows the hysteresis
YS≈1.12 belongs to the hard state for this criterion. For a clear illus- loops during the 10th cycles and half-life cycles at different strain
tration, the cyclic softening ratio (CSR) during LCF is determined as amplitudes, respectively. From Fig. 5a and b, we found increasing strain

62
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

initiated at the surface or subsurface of the samples at three different


strain amplitudes (Fig. 7). Furthermore, with increasing strain ampli-
tude (from 0.65% to 1.7%), the angle of fracture surface and stress-axis
increased in STA LSF Ti-6Al-4V samples (see inset of Fig. 7). These
findings demonstrate that strain amplitude can strongly affect the
fracture morphology of the STA LSF Ti-6Al-4V samples.
To clarify the fatigue crack propagation of the STA LSF Ti-6Al-4V
samples, LCF fractographs at two scanning velocities (900 and
1500 mm/min) and three strain amplitudes (0.65%, 1.1% and 1.7%)
were also studied. Compared with the fracture morphology (Fig. 8) of
the fatigue samples, the fatigue crack propagation features are similar
at the same strain amplitudes. For example, the fatigue striations ex-
isted in the propagation stage at strain amplitudes (0.65% and 1.1%),
but were hardly found in the samples at the large strain amplitude (Δε/
2 = 1.7%). That is, the fatigue morphology of STA LSF Ti-6Al-4V
samples exhibits different features at different strain amplitudes. Sec-
ondary cracks can be observed in all the fatigued LSF Ti-6Al-4V samples
Fig. 4. Hysteresis loops of the STA LSF Ti-6Al-4V samples under different strain (see Fig. 8). In addition, the secondary cracks increased and broadened
amplitudes. in the fracture surface at larger strain amplitudes (from 1.1% to 1.7%).
In addition, fracture morphology of a fatigued V1500 specimen at a
amplitude results in increasing area of the hysteresis loops. At the half strain amplitude of 0.65% was observed (Fig. 9). We found the fatigue
life cycles, the max stresses decreased compared with those at the 10th crack initiated from the surface pore defect of a specimen (Fig. 9a, b). In
cycles (Fig. 5a). From the Ramberg-Osgood relation [39], the cyclic the fracture surface, typical striations, secondary cracks and propaga-
stress-strain curves can be expressed as follows: tion steps were also found (Fig. 9c, d).
Moreover, we observed each fatigued surface fractograph of V900
Δε Δσ Δσ
= +( )1/0.05 (10th cycles) and V1500 specimens by SEM and the observed defects (see Fig. 10a–g)
2 2 × 110000 2 × 1080 (7)
and their sizes are listed in Table 2. Note that V1500 specimens typi-
Δε Δσ Δσ cally have more and larger defects than V900 specimens. The equiva-
= +( )1/0.05 (half life cycles) lent diameter by the square root of the area of the largest defect (i.e.,
2 2 × 100000 2 × 960 (8)
lack-of-fusion) was observed to be ∼400 μm in the V1500 specimen
To understand the detailed fatigue failure behavior of STA LSF Ti- (see Fig. 10g). These findings demonstrate that a small surface defect
6Al-4V, we also investigated the stress and plastic strain amplitudes (diameter < 60 μm) will slightly decrease the LCF lives of LSF speci-
under different strain amplitudes (Fig. 6). The stress amplitude de- mens. A similar result has been reported by Åkerfeldt et al. [11] for
creased and plastic strain amplitude increased with an increase of the Nf lack-of-defect affected LCF life in laser metal wire deposited Ti-6Al-4V.
at a given strain amplitude. These results also demonstrate cyclic soft- This result may be caused by the different defect sensitivities for various
ening behavior of STA LSF Ti-6Al-4V parts under cyclic loading con- microstructures and the small experimental error during fatigue tests.
ditions. Note that, at large strain amplitudes (Δε/2 ≥ 1.1%), the plastic
strain amplitude increased larger than that at low strain amplitudes 4. Discussions
(from 0.65% to 0.8%). The large increment of the plastic strain am-
plitude also demonstrated severe fatigue damage during the cyclic de- 4.1. Comparison of LCF properties of various AM Ti-6Al-4V samples
formation process.
In general, the LCF properties of AM Ti-6Al-4V parts are strongly
3.3. Fractographic analysis of the low cycle fatigued Ti-6Al-4V dependent on their microstructures and the presence of defects (e.g.,
pores and lack-of-fusion) in deposited parts [21]. We collected the re-
To understand the fatigue crack initiation sites, LCF fractographs at cent LCF data of AM Ti-6Al-4V (see Table 3) from refs.[8–10,40,43–45]
two scanning velocities (900 and 1500 mm/min) and three strain am- to summarize the relationships between microstructural morphology,
plitudes (0.65%, 1.1% and 1.7%) were investigated. The fatigue crack deposition parameters and LCF properties. We investigated the LCF

Fig. 5. Hysteresis loops of the STA LSF Ti-6Al-4V samples under different strain amplitudes: (a) 10th cycles, (b) Nf/2 cycles. The tips of stable hysteresis loops provide
the cyclic stress-strain curve, which are the fitting curves in (a) and (b).

63
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 6. Max stress and plastic strain amplitude versus number of cycles to failure at various strain amplitudes. (a) Max stress and number of cycles to failure (Nf)
under different strain amplitudes at a scanning velocity of 900 mm/min, (b) Max stress and Nf under different total strain amplitude at V1500; (c) Plastic strain
amplitude and Nf under different total strain amplitude at V900, and (d) Plastic strain amplitude and Nf under different total strain amplitude at V1500.

Fig. 7. LCF fractographs of STA LSF Ti-6Al-4V samples at different scanning velocities and strain amplitudes. The dash circle represents the crack origin under low
cycle fatigue processes. V900 and V1500 represent the scanning velocity 900 and 1500 mm/min, while the strain amplitude is 0.65%, 1.1% and 1.7%, respectively.
The inset is another view for the samples.

properties of LSF STA Ti-6Al-4V specimens with different prior-β grain amplitudes (from 0.55% to 1.7%) (Fig. 1b). We used the Δε/2 − 2Nf
sizes at two scanning velocities (900 and 1500 mm/min). The LCF curves to compare the differences in the LCF properties (Fig. 11a,b).
property data presented a scatter distribution at different strain These were fit by the Coffin-Manson and Basquin’s equations [39],

64
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 8. Low cycle fatigued fractographs of the STA LSF Ti-6Al-4V samples at different strain amplitudes and scanning velocities. (a) V900-0.65% represents the
scanning velocity 900 mm/min and strain amplitude 0.65%, (b) V1500-0.65%, (c) V900-1.1%, (d) V1500-1.1%, (e) V900-1.7%, and (f) V1500-1.7%.

which are the most popular equations for LCF properties. According to parts have higher LCF lives than those of other AM Ti-6Al-4V in the
the hysteresis loop at different strain amplitudes, we found that the literature, which are comparable to those of their wrought counterparts
strain amplitude from 0.55% to 1.7% has a proper parameter for the at intermediate strain amplitudes (from 0.8% to 1.1%). Ti-6Al-4V
fatigue test. For example, the higher strain amplitude (Δε/2 > 1.7%) wrought and Ti-6Al-4V extra low interstitial (ELI) wrought have su-
caused the typical fatigue life to be less than 100 cycles (see Fig. 1b), perior LCF lives than that of AM Ti-6Al-4V owing to their fine equiaxed
whereas at a lower strain amplitude (Δε/2 < 0.55%), the fatigue life α grains resulting in a combination of high strength and high ductility,
was typically greater than 104 cycles (see Fig. 1b). To distinguish this especially the Ti-6Al-4V ELI wrought [45]. Note that the differences in
slight difference of LCF curves for AM Ti-6Al-4V in Fig. 11a, we mag- LCF properties between AM Ti-6Al-4V and the wrought standard are
nified the Δε/2-2Nf curves (Fig. 11b). Thus, almost all the LCF prop- much greater with an increase in the strain amplitude (Δε/2 ≥ 1.1%)
erties of Ti-6Al-4V fabricated by AM are lower than those of their (Fig. 11a, b). Generally, the low cycle fatigue life for a given materials
wrought counterparts. (Fig. 11a, b). Note that the STA LSF Ti-6Al-4V and strain amplitude still exhibit large scatter [39]. In order to

65
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 9. Fracture morphology of a low cycle fatigued V1500 specimen at a strain amplitude of 0.65%. (a) Macro-morphology of a fatigued specimen. FCG denotes
fatigue crack propagation, (b) Fatigue crack initiated from surface pore, (c) Typical striations presented in FCG region, and (d) Fracture morphology showing
secondary cracks, striations and steps in the transition region.

eliminate the influence of the scatter of fatigue data, we use a unified properties of V1500 samples are slightly lower than those of the V900
formula (i.e. Coffin-Manson equation) for comparative study. In addi- samples. However, when 2Nf are higher than ∼5000 cycles, the LCF
tion, the fatigue data fitted by Coffin-Manson equation represents the properties of V1500 are higher than those of V900. These results de-
best confidence for each study. Thus, the substantial distinction from monstrate that increasing the scanning velocity may enhance the LCF
the Coffin-Manson equation can basically reflect the difference between properties of the LSF Ti-6Al-4V alloys at the lower strain amplitudes.
all these datasets. These results demonstrate that the LSF Ti-6Al-4V Note that the LCF lives of the V1500 specimen are slightly higher than
fabricated by high-power and certain deposition parameters have better those of the V900 specimens at low strain amplitudes regimes. This
LCF lives than other AM Ti-6Al-4V under low-power (≤4000 W) con- result may be caused by the V1500 specimens with the finer α-phase
dition from references. (5 ± 1.0 μm) and finer prior-β grains (630 ± 150 μm) than those of
In addition, the ductility of LSF parts plays a more important role in the V900 specimen with the coarser α-phase (9 ± 1.0 μm) and coarser
LCF properties than the tensile strength [39]. According to Refs. [20] prior-β grains (1000 ± 150 μm), these microstructures can be found in
and our previous work [43], we found that as-built Ti-6Al-4V parts our previous work [25]. The difference between the LCF curves for
typically have high yield strength (∼830–1000 MPa) but low ductility V900 and V1500 specimens is slightly smaller due to the small differ-
(∼7%–11%). Because the low ductility of as-built Ti-6Al-4V parts ty- ence between their tensile properties; particularly the ductility was
pically result in inferior LCF lives, the LCF properties of as-built Ti-6Al- approximately equal.
4V V900 and V1500 specimens were not reported in this work, but can Generally, the strain-life curves are often called “low cycle fatigue”
be found in our previous work [43] using similar deposition para- due to their data resulting in fewer than ∼105 cycles [46]; therefore,
meters. The difference between our previous and present work lies in we used this number (∼105 cycles) as a boundary for low and high
the scanning velocity, for example, ∼700 mm/min as reported in Ref. cycle fatigue lives. Fig. 11d shows a schematic of the relationship
[43] but 900 and 1500 mm/min in this work. Furthermore, the LCF among yield strength (YS), elongation to failure (δ) and LCF properties.
properties of LSF Ti-6Al-4V under as-deposited, STA, and HIP condi- According to the Coffin-Manson equation, (Eq. (4)), it was found that
tions were also reported (see curves F8-F10 in Table 2). We found that YS generally controlled the fatigue strength coefficient (σf′) and the
the V900 and V1500 specimens under STA condition have superior LCF fatigue strength exponent (b), whereas elongation to failure (δ) con-
lives as compared to those of our previous work under as-deposited, trolled the fatigue ductility coefficient (εf′) and the fatigue ductility
STA, and HIP conditions. exponent (c). Therefore, the improvement in YS will increase the LCF
Furthermore, the Coffin-Manson curves (Fig. 11c) from the fatigue life at low strain amplitudes, while the improvement of the δ will in-
data at two scanning velocities (900 and 1500 mm/min) were studied creases the LCF life at high strain amplitudes. Both YS and δ of AM Ti-
to evaluate the LCF properties. The LCF life curves for V900 and V1500 6Al-4V parts increase, and the LCF life will be enhanced at all strain
specimens present crossing characters within their strain amplitude amplitudes. From the influence of yield strength and ductility on LCF
conditions (Δε/2 from 0.55% to 1.7%). For example, when the number lives (see Fig. 11d), the LCF lives of V900 specimens should be higher
of reversals to failure (2Nf) is lower than ∼5000 cycles, the LCF than those of the V1500 specimens at a large strain amplitude regime

66
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 10. Fractographs of low cycle fatigued Ti-6Al-4V specimens showing the observed pores and lack-of-fusion defects. (a) V900-0.55% and Nf = 16136, (b) V1500-
0.7% and Nf = 3240, (c) V1500-1.4% and Nf = 95, (d) V1500-0.8% and Nf = 1660, (e) V1500-1.1% and Nf = 550, (f) V900-0.65% and Nf = 4505, (g) V1500-1.4%
and Nf = 194, (h) V1500-0.8% and Nf = 1660, (i) V1500-1.1% and Nf = 456. Note that the equivalent diameter area of lack-of-fusion defect (∼540 μm × 295 μm)
is ∼400 μm in (g).

(Δε/2 ≥ 1.0%). low E conditions [25,28]. Thus, for high power LSF processes, we
Fig. 12 shows a schematic of the relationship between linear energy should decrease the linear energy density to enhance the strength of AM
density and tensile properties and microstructures for the LSF process. Ti-6Al-4V parts. This result is mainly caused by the post-LSF heat
Owing to the strength-ductility trade off dilemma, we should balance treatment, which essentially sacrificed strength to increase ductility.
the strength and ductility for LSF Ti-6Al-4V parts under low and high Based on the relationship between the tensile properties and low
power LSF processes. The AM Ti-6Al-4V parts under large linear energy cycle fatigue properties, it can be seen that the LSF Ti-6Al-4V alloy has
density (E) conditions usually have lower tensile strength than under high plasticity typically has a high LCF life. Fig. 13 is a schematic of the

Table 2
The observed defects in fractographs of fatigued STA LSF Ti-6Al-4V samples.
No. Sample condition Diameter of defect size (μm) Position of defects Nf (cycles)

N1 V900-3#-0.55% 115 Surface 16,136


N2 V900-17#-0.65% 40 Interior 4505
N3 V900-12#-0.8% 35 Interior 2650
N4 V1500-39#-0.65% 150 Surface 4298
N5 V1500-40#-0.65% 190 Surface 4872
N6 V1500-26#-0.7% 105, 105, 140 Surface, others interior 3240
N7 V1500-23#-0.8% 130, 230 Interior 1800
N8 V1500-13#-0.8% 200 Surface 1660
N9 V1500-25#-1.1% 280 Subsurface 550
N10 V1500-9#-1.1% 80 Interior 669
N11 V1500-15#-1.1% 70 Interior 456
N12 V1500-38#-1.4% 35, 150, 150, 400 Interior 194
N13 V1500-21#-1.4% 290 Surface 95

Note that the defect size of 105 μm with bold-fonts is the surface defect of the sample. other defects (105, 140 μm) are the interior defects.

67
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Table 3
Low cycle fatigue Δε/2 − 2Nf curves of the Ti-6Al-4V fabricated by AM process and wrought counterparts from various references.
No Condition β grain size (μm) α morphology and average size YS (MPa) El (%) Δε/2 − 2Nf equation Ref.
(μm)

F1 LSF TC4 STA 1000 ± 150 Basket-weave α + AMellar 791.6 ± 12 18.2 ± 0.2 Δε/2 = 0.0097(2Nf)−0.05217 + 0.20621(2Nf)−0.57527 This
V900 (αs + β), 9 ± 1.0 work
F2 LSF TC4 STA 630 ± 150 Basket-weave α + AMellar 839.5 ± 11 17.8 ± 0.5 Δε/2 = 0.00946(2Nf)−0.04474 + 0.21597(2Nf)−0.60018
V1500 (αs + β), 5 ± 1.0
F3 As-built LENS ∼138 α laths, (138 μm long, 4 μm width) 908 3.8 Δε/2 = 0.022(2Nf)−0.135 + 0.030(2Nf)−0.530 [8]
F4 HT LENS ∼138 α laths, (74 μm long, 11 μm width) 959 3.7 Δε/2 = 0.015(2Nf)−0.111 + 0.736(2Nf)−0.967
F5 Annealed LENS ∼138 Widamanstatten α-laths (coarse) 957 34 Δε/2 = 0.040(2Nf)−0.210
F6 LENS 1 Large Basket-weave α (coarser) + αs 908 15.2 Δε/2 = 0.019412(2Nf)−0.136 + 0.256(2Nf)−0.797 [9]
F7 LENS 2 Small Basket-weave α 956 13.4 Δε/2 = 0.019583(2Nf)−0.127 + 7.55(2Nf)−1.290
F8 As-built TC4 200–1000 Widamanstatten α, 1–2 893 11 Δε/2 = 0.01177(2Nf)−0.07162 + 2.13535(2Nf)−1.00070 [43]
F9 STA TC4 200–1000 Basket-weave α, 2.5–3.5 896 10 Δε/2 = 0.01058(2Nf)−0.06089 + 0.77732 (2Nf)−0.83707
F10 HIP TC4 200–1000 Basket-weave α, 2.5–3 872 12.3 Δε/2 = 0.01028(2Nf)−0.05750 + 0.58990(2Nf)−0.78261
F11 LSF STA 200–500 Basket-weave α, 2–5 1020 14 Δε/2 = 0.013975(2Nf)−0.09854 + 0.703526(2Nf)−0.89922 [10]
F12 TC4ELI wrought α grain: ∼4um Equiaxed α 992 15 Δε/2 = 0.01242(2Nf)−0.060 + 4.83(2Nf)−0.898 [45]
F13 TC4 wrought Equiaxed grains Equiaxed α ≥825 ≥10 Δε/2 = 0.013 (2Nf)−0.07 + 2.69(2Nf)−0.96 [40]
F14 SLM TC4 ELI Fine acicular a’ martensite 1015 10 Δε/2 = 0.02761 (2Nf)−0.186 + 15.35(2Nf)−1.47 [44]

Note: LSF: laser solid forming, STA: solution treatment and aging, LENS, laser engineered net shape, HT: heat treated, HIP: heat isostatic pressing, ELI: extra low
interstitial, YS: yield strength, El: elongation to failure. TC4: Ti-6Al-4V.

Fig. 11. Comparison of LCF properties of AM Ti-6Al-4V and the wrought standard: (a) various AM Ti-6Al-4V samples and the wrought standard. (b) Magnification of
(a) shown in the dashed rectangle; TC4 is short for Ti-6Al-4V. (c) LCF properties of LSF Ti-6Al-4V at two scanning velocities, (d) Schematic of the relationships among
yield strength (YS), elongation to failure (δ) and LCF properties. Note: F1-F2[This work], F3-F5[8], F6-F7[9], F8-F10[43], F11[10], F12[45], F13[40], and F14[44].

relationships between LCF properties, tensile properties, and micro- sample and the α-phases. In the acceptable region, the microstructure is
structure characteristics of AM Ti-6Al-4V. In Fig. 13, the wrought usually a mixture of equiaxed or α-laths phase and fine lamellar
standard (i.e., elongation is 10%, yield strength is 825 MPa) is used as (αs + β). From the results of the LCF properties comparison, in order to
the boundary. When the tensile property is higher than the wrought obtain excellent LCF performance, the corresponding data points should
standard, this tensile property is basically qualified, as shown in the first fall into the acceptable area, and secondly in the lower right part of
acceptable region. Outside this acceptable zone, it is an unacceptable the acceptable area (more excellent plasticity). Combined with tensile
(i.e., not acceptable) zone. In the upper half of the unacceptable zone, properties and LCF properties, it can be seen that the high power LCF
the microstructure usually has a martensitic α’ phase, which reduced to Ti-6Al-4V alloy generally has a higher elongation and thus has a better
AM parts have high strength and low elongation. In the lower half of the LCF property than the AM process under low power conditions. In ad-
unacceptable zone, AM parts usually have both low strength and low dition, AM Ti-6Al-4V parts with lamellar (αs + β) microstructures ty-
elongation due to the presence of a large number of defects in the pically exhibit superior ductility, and these findings are supported by

68
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

properties. It can be found from Table 4 that a medium (∼190 μm)


surface defect dramatically decreased the LCF properties. Surprisingly,
the fatigued specimens with a small (∼60 μm) surface defect still have
a high LCF life (see Table 4) at a given strain amplitude.
To explain the effect of surface defects on fatigue life, we in-
vestigated the relationship between the defect diameter versus strain
amplitude and maximum stress amplitude for LSF Ti-6Al-4V subjected
to LCF loading. The main formulas are as follows:

Δσ ΔK
=
2 2Fw π area (9)

ΔK Δε
= σ0 [1 − exp(−A )]
2Fw π area 2 (10)

where area is Murakami’s parameter [51] for expressing the defect


size, Fw is a geometrical factor and depends on the defect location
(Fw = 0.65 for surface defect and Fw = 0.5 for internal defect), ΔK is the
stress intensity factor range, Δσ is the applied stress range. σ0 and A are
constants. Detailed descriptions of Eqs. (9) and (10) see supplementary
material A1.
Fig. 12. Schematic of microstructures and tensile properties versus linear en-
Because ΔK is unknown in this formula, and this value directly
ergy density for LSF process, Note that dβ and dα represent the prior-β grain
determines the position of the black solid line and the red dashed line, it
width and α-lath width, respectively.
corresponds to the relationship between the defect size and the stress
amplitude or strain amplitude, thus, we needs to first determine the ΔK
value. In order to obtain the effect of surface defects on fatigue life, we
choose the experimental results of the stress amplitudes of 6 surface
defects from Table 4 (see the data labeled surface in Table 4 for details),
due to the experimental point (0.55%–115 μm) is the point at which the
fatigue fracture passes through the smallest point of the surface defect
size, and the experimental point (1.1%–60 μm) is the point at which the
surface defect size has no influence on the fatigue crack initiation po-
sition. In this work, these two points are taken as the critical point for
determining ΔK, so that ΔK = 15.5 MPa m1/2 was obtained. Therefore,
the relationship between the defect and the stress amplitude and the
relationship between the defect and the strain amplitude during low-
cycle fatigue can be expressed as:

Δσ 15.5
=
2 2Fw π area (11)

15.5 Δε
= 986.13346[1 − exp( −1.99124 )]
2Fw π area 2 (12)

Fig. 13. Relationships between LCF, tensile strength and microstructure fea- Thus, it can be judged that the presence of the maximum defect size
tures for AM Ti-6Al-4V. does not affect its LCF life under a given strain amplitude.
The sample with surface defect (black curve) is more harmful
the Ref. [47]. (Fw(surface)/ Fw(interior) = 0.65/0.5 = 1.3 times) than that with interior
defect (red dash curve) to fatigue lives. The experimental data (e.g.,
from Exp. 1 to Exp. 6 in Fig. 15) were obtained by SEM observing the
4.2. Surface defects versus LCF properties fatigued fracture surface of STA LSF Ti-6Al-4V samples. Meanwhile,
from Fig. 15 the effect of the size of the defect on LCF life can be easily
For defect and fatigue properties, the well-known relation was de- predicted. If the combination of defect size and cyclic load (for ex-
scribed by the so-called Kitagawa-Takahashi diagram [48]. More re- ample, maximum stress amplitude or strain amplitude) is located in the
cently, Beretta and Romano [49] reviewed the fatigue strength of AM lower left area (Olive area) of Fig. 15, the defect may slightly reduce the
Ti-6Al-4V, especially in terms of sensitivity to defects and in- LCF life of the Ti-6Al-4V sample; if this combination is in the area above
homogeneities. M. Benedetti et al. [44] investigated the influence of the red dashed curve, the defect will significantly reduce the low cycle
mean stress and defect sensitivity on low and high cycle fatigue re- fatigue life. In addition, Fig. 15(b) can predict the effect of surface
sistance of Ti-6Al-4V ELI additively manufactured via selective laser defects and internal defects on fatigue life under different strain am-
melting process. They found the defectiveness in high cycle fatigue plitudes. For example, at a given strain amplitude, a horizontal line can
regime is well represented by the Murakami model. Tammas-Williams be drawn to intersect the black solid line and the red dashed line to
et al. [50] believed that the size and location of these defects is crucial obtain a critical value of the defect size. If the actual defect size is larger
in determining the fatigue life of AM Ti-6Al-4V samples. However, little than the critical value, the defect can be basically predicted the strain
work was reported about the surface defect and LCF properties for AM amplitude is dangerous, that is, it will greatly reduce the final LCF life;
parts. Thus, the LCF lives of LSF Ti-6Al-4V specimens at certain strain if the actual defect is less than the critical value, it can be predicted that
amplitudes (0.65% and 1.1%) were presented (see Fig. 14a–h and the defect is safe under this strain amplitude, that is, it is not significant
Table 4) to investigate the influence of surface defects on LCF reduce the LCF life.

69
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 14. Surface defects and fatigue crack initiation of the fatigued V1500 specimen: (a) marked area contains a defect (pore) at the surface of the fatigued sample, (b)
magnified defect morphology showing this ∼60 μm defect position without affect the fatigue crack initiation, (c) fatigued sample after 669 cycles, (d) typical
fractograph of the sample, (e) magnified image of the (d) showing dashed rectangle, (f) fatigue crack initiated at ∼190 μm surface defect, (g) fatigued sample after
4872 cycles, (h) fatigued surface of the specimen (Nf = 4872 cycles) corresponds to (g).

Table 4 formed by a particle or pore, Nmsc represents the number of cycles re-
LCF properties of LSF Ti-6Al-4V V1500 with surface defects at different strain quired for growth of an MSC, Npsc is the number of cycles required for
amplitudes. propagation of a PSC with length in the interval and Nlc is the number
Δε/2 (%) Nf (cycles) Surface defect size Crack initiation of cycles required for long crack growth to final failure, which depends
(μm) position on the loading level and the extent of local plasticity ahead of the crack
tip. In general, for AM Ti-6Al-4V specimens, Nlc was not observed and
0.65 4872 ∼190 Cross the defect
was thus neglected [53].
0.65 4298, 6068
1.1 669 ∼60 Not cross the defect Crack incubation at pores or inclusions was modeled by a modified
1.1 374, 456, 550, Coffin-Manson relation, as shown by Eqs. (14)–(18):
702 α
Cinc Ninc =β (14)

1 l
4.3. A Fatigue life prediction model and calibration Cinc = Cn + < − 0.3 > (Cm − Cn )
0.7 D (15)

A microstructure-based multistage fatigue (MSF) model was first P∗


Δγmax l
β= = Y (εa − εth )q , < ηlim
developed by McDowell et al. [52] for a cast A356-T6 aluminum alloy 2 D (16a)
and was calibrated by Torries et al. [53] for laser additive manu-
P∗
facturing Ti-6Al-4V. A brief recast of the MSF model is described in the Δγmax l l
β= = Y (1 + ξ )(εa − εth )q , > ηlim
following accompanied by a detailed explanation of model parameters 2 D D (16b)
for LSF Ti-6Al-4V. In the MSF model, the stages of fatigue damage D0 d′
evolution typically contain crack incubation, microstructurally small Y = [y1 + (1 + R) y2 ]( )
D (17)
crack (MSC) and physically small crack (PSC) growth as well as long
crack (LC) growth. Thus, the total fatigue life, Nt, may be considered as: l < ε − εth > l
= ηlim a , < ηlim
D εper − εth D (18a)
Nt = Ninc + Nmsc/psc + Nlc (13)
l εper r l
= 1 − (1 − ηlim )( ), > ηlim
where Ninc is the number of cycles to incubate a crack at a micronotch D εa D (18b)

70
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Table 5
List of parameters used in the MSF model for LSF Ti-6Al-4V alloy.
Definition Parameter value Ref.

Incubation
Coefficient for nucleation in HCF region Cn 0.009
Coffin-Manson coefficient for incubation Cm 0.21
Ductility exponent in Coffin-Manson law α V900: −0.575
V1500: −0.6
Exponent in remote strain to local plastic q 2.45 [53]
shear strain
Constant in remote strain to local plastic y1 100
shear strain
Shape constant for the transition to limit r 0.2
plasticity
Percolation limits from constrained to ηlim 0.3 [53]
unconstrained
Cyclic plastic strain percolation limits εper 0.52%
Fatigue damage threshold εth 0.22%

MSC/PSC growth
Constant factor to model combined stress θ 0
state effects
Material constant χ 0.32 [55]
Coefficient for the LCF regime CI 1.3 × 105
Coefficient for the HCF regime CII 0.35
Exponents for the HCF regime n 4.3 [57]
Average void volume fraction (porosity) f(ϕ) ∼1
Burgers vector for HCP Ti, b ΔCTDth 2.95 × 10-4 μm [3]

Note that the parameter without cited reference is from this work.

amplitude in the HCF regime and the range of macroscopic plastic shear
strain in the LCF regime, CI is the LCF coefficient, CII and n are the
coefficient and exponents for the HCF regime, re-

spectively.Δσ = 2θσ¯a + (1 − θ)Δσ1 is the equivalent applied stress
' '
3 Δσij Δσij
range, σ¯a = [ 2 2 2 ]0.5 is the von Mises uniaxial effective stress am-
plitude, and Δσ1 is the maximum principal stress range. The parameter θ
Fig. 15. Defects size and fatigue crack initiation of the fatigued specimen: (a) is in the range of 0≤ < θ ≤ 1, with θ = 1 as the von Mises stress
relationship between the defects size and max stress amplitudes, and (b) re- dominant case and θ = 0 as the maximum principal stress dominant
lationship between the defects size and strain amplitudes. case. Note that χ is a constant for a given microstructure, typically less
than unity, and ΔCTD is the range of cyclic crack tip displacement. DCS
where Cn is the coefficient in the HCF regime for nucleation and small is the dendrite cell size, and DCS0 is the reference dendrite cell size. The
crack growth at inclusions, α is the ductility exponent in the Coffin- function f (ϕ¯ ) represents the average void volume fraction (porosity) ϕ̄ .
Manson law, R is the load ratio, β is the nonlocal damage parameter The parameter U is the load ratio parameter in the form of U = 1/
based on the ratio between plastic zone sizes around a defect, D is the (1 − R) for R ≤ 0 and U = 1 for R > 0. A more detailed description
size of the pertinent inclusion at which fatigue incubates, l/D is the and assumption for the MSF model can be found in Refs. [54–56].
nominal ratio between the plastic zone size and the diameter of the A summary of the MSF model parameters for the LSF Ti-6Al-4V
defect that forms the fatigue damage, εth = 0.29Sut/(E(1 − R)) is the alloy is shown in Table 5. Among these parameters, the coefficients of
fatigue damage threshold, εper = (0.7 − 0.9) σycyc/ E is the cyclic plastic Eqs. (19)–(21) and previous Eqs. (14)–(18) for predicting incubation
strain percolation limit, q is the exponent in remote strain to local life were determined by Refs. [54,55,57] and this work. Once the
plastic shear strain, and ξ is the geometric factor determined in mi- parameters were determined, the incubation life, MSC/PSC life, MSF
cromechanical simulations. The r is a shape constant for the transition prediction life, as well as MSF upper and lower bounds can be obtained
to limit plasticity, and ηlim is the limiting factor that defines the tran- according to MSF model. MSF model parameters were obtained using
sition from constrained microplasticity to unconstrained microplasticity Microsoft Excel and Origin software for model correlation and optimi-
at the notch. D0 is the diameter of the pore serving as the crack in- zation for each stage of fatigue life. Note that the combination of the
itiation site, and d’ is the effect of the pore size on local plastic strain; y1 parameters of MSF model controls the prediction effect of fatigue life.
and y2 are constants, the parameter Y is correlated with y1 and y2. Fig. 16 shows the fatigue life predictions using the MSF model for
Function < f > = f if f ≥ 0; < f > = 0 otherwise. STA LSF Ti-6Al-4V alloys under V900 and V1500 conditions. Fig. 16(a,
The MSC/PSC growth law is given by Eqs. (19)–(21): c) present the contribution of crack incubation and MSC/PSC growth to
the total fatigue life, given by the MSF model, for the V900 and V1500
da
( )msc/psc = χ (ΔCTD − ΔCTDth) specimens. Fig. 16(b, d) presents the fatigue upper and lower bounds
dN (19)
for the V900 and V1500 specimens, respectively. As shown in Fig. 16(a,
∧ P
DCS U Δσ n DCS Δγmax 2 c), the fatigue life of STA LSF Ti-6Al-4V appears to be almost entirely
ΔCTD = f (φ¯)CII ( )[ ] ai + CI ( )( ) driven by the MSC/PSC growth stage, which is the same result found in
DCS0 Sut DCS0 2 (20)
Ref. [53] reporting the MSF model for AM Ti-6Al-4V. The MSF model
ϕ¯ can give satisfactory fatigue life predictions for STA LSF Ti-6Al-4V ex-
f (ϕ¯ ) = 1 + ω {1 − exp(− )}
2ϕth (21) cept for a slightly differences at low (0.55%) strain amplitudes in V900
specimens. The prediction fatigue life is slightly lower than the ex-
where ΔCTD is the crack tip displacement range, which is proportional perimental fatigue data at 0.55%, this difference may be caused by the
to crack length ai = 0.5625D, σan is the nth power of the applied stress

71
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Fig. 16. Fatigue life predictions using MSF model for STA LSF Ti-6Al-4V alloys under V900 and V1500 conditions. (a) Incubation and MSC/PSC life for V900
specimens, (b) MSF lower bound and upper bound for V900 specimens, (c) Incubation and MSC/PSC life for V1500 specimens, (d) MSF lower bound and upper bound
for V1500 specimens. The upper and lower bounds were obtained for the smallest (∼5 μm) and largest (∼400 μm) average incubation pores or inclusions. The MSF
prediction life was obtained for intermediate (∼100 μm) pores or inclusions.

parameters used in the MSF model and the large dispersity of experi- the LCF lives of STA LSF Ti-6Al-4V parts even at a high (∼1.1%)
mental data. In addition, this MSF prediction error has a similar ten- strain amplitude.
dency with that of Ref. [53], which presents large error at low strain (4) A microstructure-based multistage fatigue (MSF) model was used to
amplitude. Thus, this MSF prediction error is acceptable in this in- predict the LCF life for STA LSF Ti-6Al-4V alloy. The fatigue life
vestigation. For the lower bounds and MSF prediction life, we used the prediction obtained with this MSF model shows good agreement
largest (∼400 μm) and moderate (∼100 μm) average incubation pores with our experimental data.
or inclusions. The MSF lower bound and prediction total life are fit well
with the experimental data; this may be caused by the presence of Acknowledgments
several pores or lack-of-fusion defects in the fatigued specimens
showing similar weight factor with such large (∼400 and ∼100 μm, This work was supported by National Key Research and
respectively) pores in the MSF model. Thus, through the MSF model, Development Plan of China (2016YFB1100104).
fatigue life predictions of LSF Ti-6Al-4V can be more accurately pre-
dicted, resulting in more trustworthy LSF parts and quicker adoption of Appendix A. Supplementary material
LSF processes for industry applications.
Supplementary data to this article can be found online at https://
5. Conclusions doi.org/10.1016/j.ijfatigue.2019.05.035.

Low cycle fatigue (LCF) properties of Ti-6Al-4V alloy fabricated by References


high-power (7600 W) laser solid forming (LSF) additive manufacturing
at solution treatment and aging (STA) condition were investigated. The [1] DebRoy T, Wei HL, Zuback JS, Mukherjee T, Elmer JW, Milewski JO, Beese AM,
main findings are summarized as follows: Wilson-Heid A, De A, Zhang W. Additive manufacturing of metallic components –
process, structure and properties. Prog Mater Sci 2018;92:112–224.
[2] Gu DD, Meiners W, Wissenbach K, Poprawe R. Laser additive manufacturing of
(1) The STA LSF Ti-6Al-4V parts exhibit higher LCF lives than those of metallic components: materials, processes and mechanisms. Int Mater Rev
other AM Ti-6Al-4V in the literature, which are also comparable to 2012;57:133–64.
[3] Banerjee D, Williams JC. Perspectives on titanium science and technology. Acta
those of their wrought counterparts at intermediate strain ampli- Mater 2013;61:844–79.
tudes (from 0.8% to 1.1%). [4] Lewandowski JJ, Seifi M. Metal additive manufacturing: a review of mechanical
(2) Cyclic softening behaviors of the STA LSF Ti-6Al-4V parts are found properties. Annu Rev Mater Res 2016;46:151–86.
[5] Beese AM, Carroll BE. Review of mechanical properties of Ti-6Al-4V made by laser-
at various strain amplitudes (from 0.65% to 1.7%). With an in-
based additive manufacturing using powder feedstock. JOM 2015;68:724–34.
crease in strain amplitude, the degree of cyclic softening increased. [6] Bian L, Thompson SM, Shamsaei N. Mechanical properties and microstructural
(3) Small surface defect (diameter less than ∼60 μm) slightly decreases features of direct laser-deposited Ti-6Al-4V. JOM 2015;67:629–38.
[7] Qian M, Xu W, Brandt M, Tang HP. Additive manufacturing and postprocessing of

72
Y.M. Ren, et al. International Journal of Fatigue 127 (2019) 58–73

Ti-6Al-4V for superior mechanical properties. MRS Bull 2016;41:775–84. [32] Leuders S, Thöne M, Riemer A, Niendorf T, Tröster T, Richard HA, Maier HJ. On the
[8] Sterling A, Shamsaei N, Torries B, Thompson SM. Fatigue behaviour of additively mechanical behaviour of titanium alloy TiAl6V4 manufactured by selective laser
manufactured Ti-6Al-4V. Procedia Eng 2015;133:576–89. melting: Fatigue resistance and crack growth performance. Int J Fatigue
[9] Zhang SY. Research on the heat treated microstructures and properties of laser solid 2013;48:300–7.
forming Ti-6Al-4V alloy. In: Northwestern Polytechnical University, Xi'an; 2009. p. [33] Manson SS. Fatigue: a complex subject – some simple approximations. Exp Mech
93–96. 1965;5:193–226.
[10] Sterling AJ, Torries B, Shamsaei N, Thompson SM, Seely DW. Fatigue behavior and [34] Ricotta M. Simple expressions to estimate the Manson-Coffin curves of ductile cast
failure mechanisms of direct laser deposited Ti–6Al–4V. Mater Sci Eng, A irons. Int J Fatigue 2015;78:38–45.
2016;655:100–12. [35] Muralidharan U, Manson SS. Modified universal slopes equation for estimation of
[11] Åkerfeldt P, Pederson R, Antti M-L. A fractographic study exploring the relationship fatigue chacaracteristics. J Eng Mater Technol 1988;110:55–8.
between the low cycle fatigue and metallurgical properties of laser metal wire [36] Basquin OH. The exponential law of endurance tests. Proc ASTM, West
deposited Ti–6Al–4V. Int J Fatigue 2016;87:245–56. Conshohocken, PA 1910;10. 625–625.
[12] Sterling AJ, Torries B, Shamsaei N, Thompson SM, Seely DW. Fatigue behavior and [37] Manson SS. Behavior of materials under conditions of thermal stress. National
failure mechanisms of direct laser deposited Ti–6Al–4V. Mater Sci Eng A Advisory Commission on Aeronautics: Report 1170 Cleveland: Lewis Flight
2016;655:100–12. Propulsion Laboratory 1954. p. 317–50.
[13] Sames WJ, List FA, Pannala S, Dehoff RR, Babu SS. The metallurgy and processing [38] Coffin LF. A study of the effects of cyclic thermal stresses on a ductile metal. Trans
science of metal additive manufacturing. Int Mater Rev 2016;61:315–60. Am Soc Mech Eng 1954;76. 931–931.
[14] Gorsse S, Hutchinson C, Gouné M, Banerjee R. Additive manufacturing of metals: a [39] Sursh S. Fatigue of materials. 2nd ed., Cambridge University Press; 1998. p. 221–79.
brief review of the characteristic microstructures and properties of steels, Ti-6Al-4V [40] Aeronautical China. Materials Handbook, Vol 4: Titanium alloys and. Copper Alloys
and high-entropy alloys. Sci Technol Adv Mat 2017;18:584–610. 2001:104–32.
[15] Kasperovich G, Hausmann J. Improvement of fatigue resistance and ductility of [41] Lee Y-L, Pan J, Hathaway RB, Barkey ME. Fatigue testing and analysis: theory and
TiAl6V4 processed by selective laser melting. J Mater Process Technol practice. 2005: p. 190–219.
2015;220:202–14. [42] Shao CW, Zhang P, Liu R, Zhang ZJ, Pang JC, Zhang ZF. Low-cycle and extremely-
[16] Yu H, Li F, Wang Z, Zeng X. Fatigue performances of selective laser melted Ti-6Al- low-cycle fatigue behaviors of high-Mn austenitic TRIP/TWIP alloys: Property
4V alloy: Influence of surface finishing, hot isostatic pressing and heat treatments. evaluation, damage mechanisms and life prediction. Acta Mater 2016;103:781–95.
Int J Fatigue 2019;120:175–83. [43] Li J. Research on the microstructures and mechanical properties of high power laser
[17] Li P, Warner DH, Pegues JW, Roach MD, Shamsaei N, Phan N. Investigation of the solid formed TC4 titanium alloy. In: Northwestern Polytechnical University; 2015.
mechanisms by which hot isostatic pressing improves the fatigue performance of p. 53–58.
powder bed fused Ti-6Al-4V. Int J Fatigue 2019;120:342–52. [44] Benedetti M, Fontanari V, Bandini M, Zanini F, Carmignato S. Low- and high-cycle
[18] Liu S, Shin YC. Additive manufacturing of Ti6Al4V alloy: A review. Mater Des fatigue resistance of Ti-6Al-4V ELI additively manufactured via selective laser
2019;164. 107552–1–23. melting: Mean stress and defect sensitivity. Int J Fatigue 2018;107:96–109.
[19] Cao F, Zhang T, Ryder MA, Lados DA. A review of the fatigue properties of addi- [45] Carrion PE, Shamsaei N, Daniewicz SR, Moser RD. Fatigue behavior of Ti-6Al-4V
tively manufactured Ti-6Al-4V. JOM 2018;70:349–57. ELI including mean stress effects. Int J Fatigue 2017;99:87–100.
[20] Herzog D, Seyda V, Wycisk E, Emmelmann C. Additive manufacturing of metals. [46] Stephens RI, Fatemi A, Stephens RR, Fuchs HO. Metal fatigue in engineering. 2nd
Acta Mater 2016;117:371–92. ed. John wiley & Sons, Inc.; 2001. 93–12.
[21] Shamsaei N, Yadollahi A, Bian L, Thompson SM. An overview of Direct Laser [47] Xu W, Brandt M, Sun S, Elambasseril J, Liu Q, Latham K, Xia K, Qian M. Additive
Deposition for additive manufacturing; Part II: Mechanical behavior, process manufacturing of strong and ductile Ti–6Al–4V by selective laser melting via in situ
parameter optimization and control. Additive Manuf 2015;8:12–35. martensite decomposition. Acta Mater 2015;85:74–84.
[22] Li P, Warner DH, Fatemi A, Phan N. Critical assessment of the fatigue performance [48] Kitagawa H, Takahashi S. Applicability of fracture mechanics to very small cracks
of additively manufactured Ti–6Al–4V and perspective for future research. Int J or the cracks in the early stage. In: Proceedings of the 2nd international conference
Fatigue 2016;85:130–43. on mechanical behavior of materials, Boston. Cleveland (OH): ASM; 1976. p.
[23] Yadollahi A, Shamsaei N. Additive manufacturing of fatigue resistant materials: 627–631.
Challenges and opportunities. Int J Fatigue 2017;98:14–31. [49] Beretta S, Romano S. A comparison of fatigue strength sensitivity to defects for
[24] Huang WD, Lin X. Research progress in laser solid forming of high-performance materials manufactured by AM or traditional processes. Int J Fatigue
metallic components at the state key laboratory of solidification processing of 2017;94:178–91.
China, 3D Printing and Additive. Manufacturing 2014;1:156–65. [50] Tammas-Williams S, Withers PJ, Todd I, Prangnell PB. The influence of porosity on
[25] Ren YM, Lin X, Fu X, Tan H, Chen J, Huang WD. Microstructure and deformation fatigue crack initiation in additively manufactured titanium components. Sci Rep
behavior of Ti-6Al-4V alloy by high-power laser solid forming. Acta Mater 2017;7:1–13.
2017;132:82–95. [51] Murakami Y. Analysis of stress intensity factor of modes I, II and III for inclined
[26] Liang YJ, Wang HM. Influence of prior-β-grain size on tensile strength of a laser- surface cracks of arbitrary shape. Eng Fract Mech 1985;22:101–14.
deposited α/β titanium alloy at room and elevated temperatures. Mater Sci Eng, A [52] McDowell DL, Gall K, Horstemeyer MF, Fan J. Microstructure-based fatigue mod-
2015;622:16–20. eling of cast A356–T6 alloy. Eng Fract Mech 2003:49–80.
[27] Zhang SY, Lin X, Chen J, Huang WD. Effect of solution temperature and cooling rate [53] Torries B, Sterling AJ, Shamsaei N, Thompson SM, Daniewicz SR. Utilization of a
on microstructure and mechanical properties of laser solid forming Ti-6Al-4V alloy. microstructure sensitive fatigue model for additively manufactured Ti-6Al-4V.
Chin Opt Lett 2009;7:498–501. Rapid Prototyping J 2016;22:817–25.
[28] Yusuf SM, Gao N. Influence of energy density on metallurgy and properties in metal [54] McDowell DL, Gall K, Horstemeyer MF, Fan J. Microstructure-based fatigue mod-
additive manufacturing. Mater Sci Technol 2017;33:1269–89. eling of cast A356–T6 alloy. Eng Fract Mech 2003;70:49–80.
[29] Wu XH, Liang J, Mei J, Mitchell C, Goodwin PS, Voice W. Microstructures of laser- [55] Xue Y, McDowell DL, Horstemeyer MF, Dale MH, Jordon JB. Microstructure-based
deposited Ti–6Al–4V. Mater Des 2004;25:137–44. multistage fatigue modeling of aluminum alloy 7075–T651. Eng Fract Mech
[30] Zhai YW, Galarraga H, Lados DA. Microstructure evolution, tensile properties, and 2007;74:2810–23.
fatigue damage mechanisms in Ti-6Al-4V alloys fabricated by two additive manu- [56] Xue Y, Pascu A, Horstemeyer MF, Wang L, Wang PT. Microporosity effects on cyclic
facturing techniques. Procedia Eng 2015;114:658–66. plasticity and fatigue of LENS™-processed steel. Acta Mater 2010;58:4029–38.
[31] Kobryn PA, Moore EH, Semiatin SL. The effect of laser power and traverse speed on [57] Zhai YW, Lados DA, Brown EJ, Vigilante GN. Fatigue crack growth behavior and
microstructure, porosity, and build height in laser-deposited Ti-6Al-4V. Scr Mater microstructural mechanisms in Ti-6Al-4V manufactured by laser engineered net
2000;43:299–305. shaping. Int J Fatigue 2016;93:51–63.

73

You might also like