Room and High Temperature High-Cycle Fatigue Properties of Inconel 718 Superalloy Prepared Using Laser Directed Energy Deposition

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Materials Science & Engineering A 825 (2021) 141865

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Room and high temperature high-cycle fatigue properties of Inconel 718


superalloy prepared using laser directed energy deposition
Xiaobin Yu a, b, *, Xin Lin a, b, **, Zihong Wang a, b, Shuya Zhang a, b, Xuehao Gao a, b,
Yufeng Zhang a, b, Yongming Ren c, Weidong Huang a, 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
School of Materials and Chemical Engineering, Xi’an Technological University, Xi’an, Shaanxi, 710021, PR China

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

Keywords: The high-cycle fatigue (HCF) properties of Inconel 718 (IN718) superalloy prepared using laser directed energy
Additive manufacturing deposition (LDED) were investigated at room temperature (RT) and 650 ◦ C. A specified heat treatment scheme,
Directed energy deposition namely homogenization + solution + aging (HSA), was employed for the as-deposited LDED IN718 alloy. The
Inconel 718
results indicate that the grains of the HSA LDED IN718 alloy have a typical bimodal grain (BG) distribution
High cycle fatigue
Bimodal grain
consisting of coarse grains (CGs) and fine grains (FGs). The fatigue limit (σ w ) of the HSA IN718 alloy at RT was
Failure mode higher than that at 650 ◦ C (~385 and ~280 MPa, respectively), and their σ w values were only ~77% and ~63%
those of wrought IN718, respectively. Fractographic analysis revealed that the fatigue failure model changed
from surface slip failure to large crystallographic facet failure with increasing testing temperature. The HCF
properties of the HSA IN718 alloy with BG microstructure are strongly affected by CGs in the material, which
eclipses the deleterious effects of the initial pore defects in the deposit. Finally, the Basquin model was fitted to
extract the stress-life curves of the HSA IN718 alloy. A certain degree of deviation was observed when predicting
the σ w value of the HSA IN718 alloy at 650 ◦ C using the Murakami model.

1. Introduction Yadollahi et al. [16] reported that the HCF cracks in SLM IN718
initiate from pores or large unmelted regions located near or on the
Additive manufacturing (AM) is deemed a disruptive technology that specimen surface. Using crystal plasticity simulation, Prithivirajan et al.
enables the fabrication of high-performance complex components with a [17] observed that the critical pore size for fatigue failure is 20 μm in
topology optimization design [1]. Laser directed energy deposition SLM IN718 with an average grain size of 48 μm. Gribbin et al. [18]
(LDED), a metal AM technology with a high deposition rate, has been studied the microstructure of SLM IN718 after heat treatment based on
successfully used to manufacture many metallic parts, such as Ti-based AMS 5663 and found that the porosity and high volume fraction of δ
[2–4], Fe-based [5–8], and Ni-based [9–13] alloy parts. Inconel 718 phases significantly deteriorate the HCF properties of SLM IN718 at
(IN718) Ni-based alloys are widely used in gas turbine blades, aero room temperature (RT) but not as much at 500 ◦ C. However, coarse
engines, and other high temperature (HT) structural parts because of grains (CGs) with a high volume fraction of twins seriously deteriorate
their prominent weldability, manufacturability, and mechanical prop­ the HCF properties at 500 ◦ C. On the other hand, Wan et al. [19] indi­
erties, especially their high HT strength and excellent fatigue properties cated that the ultrafine-scale acicular δ phase can reduce the disad­
up to 650 ◦ C [14]. High-cycle fatigue (HCF) failure is a primary failure vantageous effects of the pores on the HCF performance of SLM IN718.
mode observed in IN718 parts prepared by traditional manufacturing Balachandramurthi et al. [20] claimed that the RT HCF crack propa­
methods (e.g., casting and forging) [15]. Consequently, increasing gation path is transgranular in SLM IN718 after solution and aging
attention has been paid to the HCF behavior of IN718 alloys fabricated treatment and is not influenced by NbC, TiN, or δ phases in it.
by AM methods, such as LDED and selective laser melting (SLM). Sui et al. [21] investigated the HCF properties of LDED IN718 alloys

* Corresponding author. State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an, Shaanxi, 710072, PR China.
** Corresponding author. State Key Laboratory of Solidification Processing, Northwestern Polytechnical University, Xi’an, Shaanxi, 710072, PR China.
E-mail addresses: yuxiaobin602@163.com (X. Yu), xlin@nwpu.edu.cn (X. Lin).

https://doi.org/10.1016/j.msea.2021.141865
Received 15 May 2021; Received in revised form 29 July 2021; Accepted 4 August 2021
Available online 5 August 2021
0921-5093/© 2021 Elsevier B.V. All rights reserved.
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

at 650 ◦ C after direct aging (DA) treatment and found that unbroken produced by the plasma rotating electrode process. The chemical com­
Laves phases are beneficial for hindering fatigue crack propagation positions of these powders are shown in Table 1, and their morphologies
(FCP) at low stress amplitudes (σa ) and improve HCF life. However, and sizes (53–106 μm) are shown in Fig. 2. Optimized processing pa­
Laves phases as weakening phases are easily broken in the FCP region at rameters of LDED IN718, obtained based on our previous works [12,13],
high σa , leading to the formation of microscopic holes and cracks at are listed in Table 2. An Ar protective atmosphere (O content ≤ 50 ppm)
Laves/γ interfaces. Consequently, FCP occurred along the rupture and a cross-direction raster scanning strategy (Fig. 1b) were used during
interface, resulting in the degradation of HCF life in DA LDED IN718. In fabrication. To eliminate residual stress and develop a consistent
addition, Deng et al. [22] suggested that the residual stress in SLM IN718 microstructure in the as-deposited IN718 bulk, the HSA process was
after DA treatment is close to that in as-deposited SLM IN718, indicating adopted based on the optimized heat treatment regime used in our
that the residual stress could influence the fluctuations in the HCF data previous works [12,13]. The specific scenarios of this HSA regime are as
of the DA LDED IN718 specimen. follows: homogenization treatment (1100 ◦ C, 1.5 h/water cooling) +
It is reported that fatigue cracking in as-cast and wrought IN718 is solution treatment (980 ◦ C, 1 h/water cooling) + double aging (720 ◦ C,
initiated in Nb-rich carbides [23], casting defects [24], and favorably 8 h/furnace cooling at 50 ◦ C/h to 620 ◦ C, 8 h/air cooling) according to
oriented grains [25,26]. In terms of low-cycle fatigue, Fournier et al. the AMS 5383 and 5662 standards for casting and wrought IN718. HCF
[27] discovered that the fracture mechanism of wrought IN718 is specimens were machined from the heat-treated LDED Inconel 718 bulk,
transgranular at RT and a mixture of transgranular and intergranular at and their dimensions and orientation with respect to the deposition di­
HTs. Kawagoishi et al. [28] reported that the HCF limit of wrought rection are shown in Fig. 1c.
IN718 specimens is much higher at HTs than at RT because of the
softening of the γ matrix and surface oxidation at HTs. 2.2. Microstructure characterization
Previous research has shown that there is an obvious difference in
defects and microstructural characteristics between wrought IN718 and After the HSA specimens were etched with a solution comprising 20
SLM or LDED IN718 [1]. Furthermore, existing studies typically focus on mL HCl, 10 mL H2O2, and 10 mL H2O for ~30 s, microstructural char­
the effects of surface roughness, defects, and build orientation on the acterization was performed using optical microscopy (OM, Olympus-
fatigue properties of as-deposited AM IN718 alloys. However, AM IN718 GX71) and scanning electron microscopy (SEM, Zeiss Supra 55). The
alloys exhibit a typical bimodal grain (BG) distribution after homoge­ Vickers hardness of the HSA specimens was measured using a LECO
nization + solution + aging (HSA) heat treatment. The literature on the micro/macro automatic hardness testing system. The load and dwell
influence of BG on the HCF properties of AM IN718 alloys, particularly time were 500 g and 15 s, respectively. The grain configuration and
HT HCF, is still limited. Therefore, the purpose of this study was to crystallographic texture for the HSA specimen were obtained using an
investigate the effect of BG on the HCF properties of LDED IN718 alloys. electron backscattered diffraction (EBSD) system in a SEM (Tescan
In this study, stress-controlled HCF tests of LDED IN718 after a Mira3) under an acceleration voltage of 20 kV, a specimen tilt angle of
specified heat treatment process, i.e., HSA, were conducted under fully 70◦ , and a scan step size of 2.5 μm. The distribution of alloy elements
reversed loading at RT and 650 ◦ C. The Basquin model was applied to was measured using electron probe microanalysis (EPMA, Shimadzu
describe the stress-life (S–N) curves for the studied material, and the 1720).
Murakami model was used to predict the fatigue limit of the HSA IN718
specimen at 650 ◦ C. The available fatigue data on wrought IN718 alloys
from the literature were used for further comparison. Finally, the 2.3. Mechanical properties
assessment of fatigue fracture surfaces and longitudinal sections of fa­
tigue fractures of the tested specimens is expected to reveal the fracture Prior to HCF testing, uniaxial tensile tests were conducted to measure
mechanism and FCP behavior of HSA LDED IN718. the strength-ductility of the HSA LDED IN718 alloy using an INSTRON
3382 universal testing system under a strain rate of 1 mm/min in air at
2. Experimental procedures RT and 650 ◦ C. Tensile testing specimens with a gauge diameter of 5
mm, gauge length of 25 mm, and full length of 80 mm were machined
2.1. Material and specimen fabrication from the heat-treated LDED IN718 bulk. The orientation of the tensile
axes in the HSA specimen was identical to that in the HCF specimen. To
An Inconel 718 alloy bulk (145 × 85 × 110 mm3, Fig. 1a) was investigate the HCF performance of the HSA IN718 alloy, all specimens
fabricated in an LDED system using commercial Inconel 718 powder were tested using a high-frequency fatigue tester (QBG-50) under a load
ratio of R = − 1 with different σa values at 140 Hz in air at RT and 650 ◦ C.

Fig. 1. A LDED Inconel 718 alloy bulk (145 × 85 × 110 mm3) (a), schematic diagram of cross-direction raster scanning path pattern (b), the geometric dimensioning
and orientation of HCF specimen (c).

2
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Table 1
Chemical composition of Inconel 718 superalloy powder.
Element Ni Cr Nb Mo Al Ti Co Mn Cu

wt. % 53.24 19.24 5.19 3.03 0.51 0.98 0.32 0.065 0.046
Element Si S P B N C H O Fe
wt. % 0.077 0.001 0.009 0.004 0.015 0.025 0.006 0.001 Bal.

The tensile-compressive loading mode and sinusoidal waveform were


investigated on all specimens. The associated number of cycles to failure
(Nf ) was documented using a QBG-50 fatigue machine.
To study the fatigue failure mechanisms at RT and 650 ◦ C, fatigue
fractures and longitudinal sections of the associated fractures were
investigated using an SEM equipped with an energy dispersive X-ray
spectroscopy (EDS) system. Specimens for double Cs corrector trans­
mission electron microscopy (TEM, Themis Z) were acquired from thin
slices adjacent to the fatigue fracture regions of the tested specimens.
Thin foils were prepared by electrolytic dual-jet thinning in a solution
(HClO4: C2H5OH = 1 : 9) at − 20 ◦ C and 40 mA. TEM observations were
carried out on a Themis Z.

3. Results
Fig. 2. SEM image of Inconel 718 alloy powder.

3.1. Initial microstructure


Table 2
The main processing parameters for LDED Inconel 718 superalloy. The grain configuration of the HSA LDED IN718 alloy showed an
obvious BG (10–500 μm) distribution consisting of CGs and fine grains
Laser Scanning Beam Increment of Powder Overlap
(FGs), as shown in Fig. 3a. A similar BG microstructure for IN718 alloys
power speed (mm/ diameter Z-axis (mm) feeding (%)
(kW) min) (mm) rate (g/ fabricated using LDED technology was also reported by Lin et al. [12,13,
min) 29] and Zhu et al. [30], who believed that the inhomogeneous distri­
2.5 600 5 0.6 25 50 bution of residual stress stored in the as-fabricated LDED IN718 alloy led
to a difference in recrystallized grains after homogenization heat treat­
ment at 1100 ◦ C/1.5 h.

Fig. 3. Microstructure of the HSA IN718 alloy: (a) OM image; (b) SEM image showing the carbide, GB-δ phase and twin; (c) high magnification SEM image showing
GB-δ, γ′ ′ and γ′ phases.

3
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

In terms of the second phase of the HSA LDED IN718 alloy, detailed the HSA specimen demonstrates equiaxed grains with a high content of
SEM analyses (Fig. 3b and c) confirmed that the HSA specimen mainly annealing twins.
contained abundant γ′′ and γ′ phases (~30 nm), submicron and EPMA analyses showed that the distribution of Nb in the γ matrix was
dispersive carbides, and submicron and short rod-like δ phases precip­ homogeneous except for the enrichment in NbC precipitates (Fig. 5).
itating in grain boundaries (GBs). Because the solution heat treatment There were also a few particles rich in Ti and N, which are believed to be
temperature (980 ◦ C) employed under the HSA condition is slightly TiN phases. These TiN phases have also been discovered in LDED IN718
lower than the δ solvus temperature (~1010 ◦ C) [31], a few δ phases [33] and other Ni-based alloys [34,35]. Moreover, Fig. 5 shows that the
were mainly precipitated along the GBs (Fig. 3b). This is attributed to center of the TiN particle was rich in Al. Kim et al. [34] pointed out that
the preferential occurrence of heterogeneous δ nucleation on the GBs the nucleation of TiN particles may occur on Al2O3 particles because Al
and incoherent twin boundaries (TBs) [32], whereas intragranular in the melt is easily oxidized and forms Al2O3 nanoparticles. Neverthe­
precipitation was rather sluggish. Hence, the jagged appearance of a less, Xie et al. [31] reported that the volume fraction of TiN phases is
grain boundary (GB) is ascribed to δ precipitation at sites neighboring much lower than that of NbC phases in IN718 alloys (<2 vol%).
the GB (Fig. 3b).
EBSD technology was used to investigate the grain features and 3.2. Tensile properties
crystallographic texture of the HSA LDED IN718 alloy. The γ grains in
the HSA LDED IN718 alloy were quantitatively characterized using The static and dynamic load parts used in RT and HT environments,
orientation imaging microscopy (OIM). Fig. 4a shows an inverse pole which require a high ultimate tensile strength (UTS), yield strength (YS),
figure map, which reveals that the different colors are largely distributed and elongation at these temperatures, are generally fabricated using
evenly within the field of vision. Fig. 4b shows the recrystallized IN718 superalloys. Fig. 6a shows the representative tensile stress-strain
microstructure of the HSA specimen analyzed using the CHANNEL 5 responses measured on HSA IN718 coupons at RT and 650 ◦ C. SEM
software. The volume fractions of recrystallized, substructured, and fractographic analysis revealed that the tensile failure mode of the tested
deformed grains in the HSA specimen were ~97.6%, 2.38%, and 0.02%, specimens at both RT and 650 ◦ C was microvoid coalescence ductile
respectively. The calculated pole figures in Fig. 4c indicate that the HSA fracture. The UTS, YS, and elongation bar graphs for HSA IN718 and
specimen possesses a weak crystallographic texture (max: 3.15). In wrought IN718 (AMS 5662) are displayed in Fig. 6b. The tensile prop­
addition, the misorientation plot shows that the associated proportion of erties (UTS, YS, and elongation) of the HSA specimens were higher at RT
annealing TBs (<111>60◦ ) reached ~63 vol% (Fig. 4d), implying that than at 650 ◦ C. Fig. 6b also shows that the tensile properties of HSA

Fig. 4. OIM characterization of γ grains in the HSA IN718 alloy: (a) inverse pole figure map; (b) recrystallization microstructure; (c) pole figures; (d) grain boundary
orientation distribution map.

4
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Fig. 5. Distribution of Nb, C, Ti, N and Al element in the HSA IN718 specimen.

Fig. 6. Tensile properties of the HSA IN718 specimen at RT and 650 ◦ C: (a) the engineering stress-strain curves, and the micro-fracture morphologies of corre­
spondingly failed tensile specimens embedded; (b) comparison of the associated yield strength (YS), ultimate tensile strength (UTS) and elongation; (c) true stress-
strain (σ − ε) curves, work-hardening rate (θ) curves, and the macrographs of correspondingly failed tensile specimens embedded.

specimens at both RT and 650 ◦ C exceeded those of the wrought (AMS Fig. 6c shows the true stress-strain curves and WH rate (θ = dσ/dε vs.
5662) specimens. In particular, the HSA specimens exhibited excellent ε, where σ is the true stress and ε is the true strain) curves of the HSA
elongations of ~26.3% and ~18.2% at RT and 650 ◦ C, respectively, specimens at RT and 650 ◦ C. The WH rate curves of HSA specimens at RT
which are higher than the low limiting of elongation (12%) of the and 650 ◦ C can be generally characterized as having three stages: an
wrought IN718 alloy (AMS 5662). In conclusion, the LDED IN718 abrupt decrease in WH (stage I), a slight plateauing of WH (stage II), and
specimens after treatment with HSA still possess excellent static strength a weakening of WH (stage III). In stage I, the ε is small and the WH rate is
and plasticity at both RT (~1106 MPa, ~26.3%) and 650 ◦ C (~938 MPa, high, which is attributed to the rapid dynamic recovery caused by the
~18.2%). climb and cross-slip of dislocations [33]. However, the WH rate during
Work hardening (WH), also known as strain hardening, is the stage II mainly remains constant, which is attributable to the saturation
mechanism of strengthening a metal through plastic deformation. Dur­ of mechanical twins and dislocation multiplication [36]. As the strain
ing the deformation process, WH enhances IN718 mainly by means of increases further, in stage III, the HSA IN718 specimen does not continue
dislocation strengthening, precipitated-phase strengthening, GB to deform homogeneously, and necking occurs when dσ/dε = σ . As
strengthening, and deformation twinning strengthening. The WH shown in Fig. 6c, the necking of HSA specimens at RT and 650 ◦ C
behavior of the HSA LDED IN718 alloy can be understood by analyzing occurred at true strains of ~23.2% and ~15.1%, respectively. The entire
its stress-strain curve. fracture surface of the HSA specimen at RT was parallel to the maximum

5
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

shear stress, and the angle between the shear fracture surface and
principal stress was ~45◦ , which is similar to that of the HSA specimens σ a = σˊf (2Nf )b (1)
at 650 ◦ C.
Nevertheless, there was a significant difference between the θ curves where σa is the nominal stress amplitude (MPa), σˊf is the fatigue strength
of the HSA specimens at RT and 650 ◦ C. The WH rate of the HSA LDED coefficient (MPa), Nf is the number of cycles to failure, and b is the
IN718 alloy at RT was higher than that at 650 ◦ C, indicating that the HT Basquin exponent.
environment at 650 ◦ C restrained the climb and cross-slip of disloca­ The expressions and Basquin parameters of different specimens fitted
tions, leading to accelerated dislocation multiplication. In addition, according to the formula mentioned above and the experimental data
heterogeneous grains (10–500 μm) with annealing twins in the HSA are shown in Fig. 7 and Table 3. Although the BG microstructure led to
specimen decreased its θ at 650 ◦ C compared to that at RT, which stems the dispersion of fatigue data at both RT and 650 ◦ C to some extent, the
from the inhomogeneity of plastic deformation. Notably, the GB is a confidence coefficients (R-square) of all fitted equations exceeded 80%.
vulnerable spot in an HT environment. Hence, at 650 ◦ C, micron-sized This indicates that the Basquin equation can be suitably applied to
TBs with a high volume fraction of ~63 vol% decreased the HT WH extract the HCF data of IN718 alloys at both RT and 650 ◦ C. The σˊf value
rate and the uniform plastic deformation capacity of the HSA specimen. of the HSA specimen clearly decreased with increasing temperature
Overall, the HSA specimen still had good uniform elongation rates at from RT to 650 ◦ C, which shows good agreement with the trend of the σ w
both RT and 650 ◦ C. value for this specimen (Table 3). However, the σ ˊf value of the HSA
LDED IN718 alloy was less than that of wrought IN718 at RT but was
3.3. HCF properties greater at 650 ◦ C. In addition, the b value mainly mirrored the cumu­
lative fatigue damage rate of IN718 alloys, which is closely related to the
Fig. 7 shows the S–N curves of the HSA LDED IN718 specimen under fatigue properties. A detailed discussion is provided in section 4.
R = − 1 at RT and 650 ◦ C, where the fatigue endurance limit (σ w , Nf =
107 cycles) stresses were ~385 and ~280 MPa, respectively. The asso­
ciated HCF data for wrought IN718 alloys reported in the literatures are
also plotted for comparison. The σw values of the HSA IN718 alloy at RT Table 3
and 650 ◦ C were much lower than those of wrought IN718 alloys (~500 The values of fatigue limit and Basquin parameters in different IN718 alloys.
and ~445 MPa, respectively) [25,28,37], and their σw values were only Material group Test temperature σw (MPa) σˊf (MPa) b
~77% and ~63% those of the wrought IN718 alloy at RT and 650 ◦ C,
respectively. To characterize the fatigue behavior of the HSA IN718 HSA RT 385 8104.4 − 0.2148
Wrought [25,28] RT 500 11,274 0.2195
alloy, regression analysis was conducted to fit the S–N fatigue data with

HSA 650 ◦ C 280 4651.7 − 0.2049
the Basquin equation [38]: Wrought [37] 650 ◦ C 445 1240.1 − 0.0667

Fig. 7. Comparison of the HCF data for HSA- and wrought- IN718 alloys [25,28,37] under R = − 1 at different temperatures: (a) data collected at RT; (b) data
collected at 650 ◦ C; (c) HCF data of the HSA specimens at RT and 650 ◦ C; (d) data for all specimens tested here and data from wrought IN718 alloys.

6
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

3.4. HCF fractography also a few micropores (~2 μm) in these facets (Fig. 9f). Cracks with
river-like patterns and feather traits stemmed from these facets and
Figs. 8 and 9 show the fracture surface morphologies of the HCF propagated to the ambient zones, revealing typical cleavage fracture
specimens at RT and 650 ◦ C. The zones of the fracture surface are features (Fig. 9b, d and f). Ni-based superalloys, as face-centered-cubic
designated as the fatigue crack initiation zone (FCIZ), FCP zone (FCPZ), (FCC) crystallographic structures, easily glide along the {111}<110>
and final rupture zone (FRZ) (see Fig. 8a, c and e; Fig. 9a, c and e). At RT, slipping system. Facet rupture was prevalent in the HCF process for the
fatigue cracking of the tested specimens at different σa values was Ni-based superalloys at certain temperatures, and the large crystallo­
mainly initiated from the specimen surface or near-surface (Fig. 8b, graphic facets are deemed to be {111} planes [39]. According to the
d and f). There were obvious slip traces and a few micropores (~2 μm) results shown in Fig. 9b, d and f, the major axis length of the large
on the fracture surface of the tested specimen (Fig. 8b). In the fatigue crystallographic facets was 200–750 μm, which is in line with the major
damage process, the cyclic slip zone formed after the σa was applied to axis length of the CGs of the BG microstructure in the HSA IN718
the HSA IN718 alloy repeatedly for a long time. Consequently, the fa­ specimen (see Fig. 3a).
tigue crack could initiate from slip bands in CGs adjacent to the spec­ Fatigue striation (FS) was the most important characteristic of the
imen surface, and then propagate and coalesce with stepwise features FCPZ, as indicated in Fig. 10a and b. In the FCPZ, fatigue striations were
and river-like patterns (Fig. 8d and f). formed when cracks ran through the TB zones, which were found in the
Fig. 9 shows the SEM morphologies of the HCF fracture at 650 ◦ C. For tested specimens at both RT and 650 ◦ C (Fig. 10a and b). In addition,
HT HCF, fatigue crack initiation (FCI) at different σ a values mainly some secondary cracks along the ductile fatigue bands were visible in
stemmed from the large crystallographic facets. Furthermore, there was the tested specimen at 650 ◦ C (Fig. 10b). In the FRZ, the fracture mode of

Fig. 8. Surface failure mechanism of HSA IN718 specimens at RT: (a, b) 400 MPa, Nf = 1,069,500 cycles; (c, d) 500 MPa, Nf = 284,700 cycles; (e, f) 700 MPa, Nf =
38,100 cycles.

7
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Fig. 9. Fatigue failure mechanism of HSA IN718 specimens at 650 ◦ C: (a, b) 320 MPa, Nf = 3,476,900 cycles; (c, d) 445 MPa, Nf = 57,300 cycles; (e, f) 495 MPa, Nf
= 32,300 cycles.

the tested specimens at both RT and 650 ◦ C was very similar to that of where C is 1.43 for a surface inclusion and is 1.56 for an interior in­
uniaxial tensile rupture (Fig. 6), i.e., the fracture mode of the HCF tested √̅̅̅̅̅̅̅̅̅
clusion, HV is the Vickers hardness (kgf/mm2), and areaeff is the
specimens at both RT and 650 ◦ C was largely transgranular. In contrast,
square root of the effective area of the defect projected vertically onto
the dimples in the HCF tested specimens at RT were smaller and deeper
the fatigue load axis (μm2).
than those at 650 ◦ C (Fig. 10c and d), indicating that the HSA LDED
For HCF at 650 ◦ C, Kobayashi et al. [26] demonstrated that the facets
IN718 alloy has better ductility at RT than at 650 ◦ C. This is in accor­
behave in the same way as non-metallic inclusions. Substituting the
dance with the tensile testing result mentioned above (Fig. 6). √̅̅̅̅̅̅̅̅̅
information from Fig. 9d ( areaeff ≈ 124,943 μm2) and the Vickers
hardness value of the HSA specimen (~450 HV0.5 ) into Eq. (2), the σ w
4. Discussion
value can be calculated as 306.5 MPa, which is slightly higher than the
experimental σ w value (280 MPa). This indicates that the CGs in the HSA
4.1. Prediction of HT fatigue limit
specimen could lead to a significant fatigue limit reduction at 650 ◦ C,
which could eclipse the deleterious effect of the initial pore defects in the
To assess the influence of defects (e.g., non-metallic inclusions) on
deposit. Similarly, Kobayashi et al. [26] investigated the fatigue prop­
the fatigue limit (σ w , MPa) of the AM IN718 alloys, Murakami et al. [40,
erties of wrought IN718 alloy with different grain sizes at 550 and
41] proposed the following formula:
600 ◦ C, and found that the fatigue strength of the CGs alloy (grain size of
/ √̅̅̅̅̅̅̅̅̅
σ w = C (HV + 120) ( area eff )1/6 (2) 90 μm) decreased notably beyond 105 cycles compared to that of the FGs
alloy (grain size of 10 μm), implying that the CGs seriously deteriorate

8
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Fig. 10. Typical morphologies of propagation zone (a, b) and final rupture zone (c, d). (a, c) RT/400 MPa, Nf = 1,069,500 cycles and (b, d) 650 ◦ C/320 MPa, Nf =
3,476,900 cycles.

the HT fatigue properties of the wrought IN718 alloy. Fig. 11 shows the SF distributions of crystals before deformation with
two groups of slip systems, namely {111}<110> and {111}<112>. A
4.2. FCP path color change from blue to red, corresponding to an SF ranging from 0 to
0.5, implies a change in crystallographic orientation from hard to soft.
In the fatigue damage process, the slip system of FCC crystals is There was an obvious difference in SF distribution between CGs and FGs,
and the FGs mainly showed a soft orientation while a few CGs showed a
commonly {111}<110>. However, Stinville et al. [42] and Kong et al.
[43] reported that an additional slip system of {111}<112> is more hard orientation. Hence, unmatched deformation among the grains
occurred during HCF loading. To obtain a coordinated deformation ef­
likely to be activated at HTs. In our previous work [13], we also found
that the grains of the HSA IN718 specimen had a BG distribution. fect, the inhomogeneous deformation can be dispersed into the ambient
FGs owing to their soft orientation. Consequently, in the HT fatigue
However, different grains could possess different plastic deformation
capacities under HCF loading. Hence, the EBSD data of the HSA IN718 process, the fatigue crack is likely to initiate from the crystallographic
facets in the CGs region, as shown in Fig. 9. Moreover, strain localization
specimen from the literature [13] were further analyzed using CHAN­
NEL 5 software to investigate the Schmid factor (SF) distributions of BG and accumulation mainly occur near the micron-sized TB zone in the
CGs [42], providing a long dislocation slip distance, which is likely to
in different slip systems. The results are shown in Fig. 11.

Fig. 11. The crystal deformation ability of HSA IN718 alloy with (a) {111}<110> slip system and (b) {111}<112> slip system (the blue lines present Σ3 TBs, and
the black lines present other grain boundaries). (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of
this article.)

9
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

cause crack propagation until failure. significantly lower than that of wrought IN718 alloys, which also
To better explain the fatigue fracture behavior, Fig. 12 shows the possess a similar GB-δ phase [31]. Balachandramurthi et al. [20] also
HCF crack propagation paths at different temperatures. For RT HCF, FCP studied the HCF properties of IN718 alloys manufactured by electron
predominantly occurred through the CGs and FGs regions in a trans­ beam melting and SLM, and found that the fatigue crack path was un­
granular manner (Fig. 12a and b). Local crack branching around the affected by precipitate phases such as NbC, TiN, and δ. In addition,
main fatigue crack was also observed, which further formed secondary because the volume fraction of submicron carbides in the γ matrix of the
cracks in the associated fatigue fracture. Like the FCP path at RT, FCP at HSA specimen was only ~2 vol% [33], and the volume fraction of TiN
650 ◦ C still predominantly occurred in a transgranular manner (Fig. 12c phases was much less than that of carbides in IN718 alloys [31]. Thus,
and d), although intergranular fracture occurred occasionally. In addi­ we can speculate that these carbides and TiN phases could not seriously
tion, the fatigue crack at 650 ◦ C easily initiated and propagated along decrease the σw value and fatigue life of HSA specimens at both RT and
the TB, which implies that the TB at HT HCF loading has a weak role in 650 ◦ C, compared with wrought IN718 alloys possessing similar car­
hindering the FCI and FCP. Miao et al. [44] and Yeratapally et al. [45] bides and TiN phases.
reported a similar phenomenon, and they believed that there is a high Fig. 14 shows the presence of slip bands of the HSA IN718 specimen
concentration of elastic stress anisotropy and accumulated plastic strain after the HCF test, revealing sheared γ′′ particles. The formation of these
in the vicinity of the TB, which induces FCI and FCP along the TB. planar slip bands in IN718 alloys is because of their low stacking-fault
The phenomenon of the crack initiating twin and parent grain pair energy, which results in the formation of wider extended dislocations
that causes the specimen to fracture is found in the HCF process of the with a decrease in the capacity of screw dislocations to cross slip on
HSA IN718 specimen at 650 ◦ C. The corresponding EBSD map of the other slip planes. A group of bands located on the (111) slip plane was
crack initiating twin and parent grain pair (twin T1-Grain G1) is dis­ found in a HCF specimen fatigued at σa = 400 MPa and RT (Fig. 14a).
played in Fig. 12c. Here, the FCP direction was perpendicular to the HCF However, two groups of bands located on the (111) and (111) slip
loading direction. The {111}<110> and {111}<112> slip systems with planes, and slip was active synchronously, which were found in an HCF
the corresponding Schmid factors (SFs) for the twin and parent grain specimen fatigued at σ a = 320 MPa and 650 ◦ C (Fig. 14d). Hence, the
pair were calculated, as shown in Tables 4 and 5. The T1-G1 grain pair cyclic deformation of the HSA IN718 specimen fatigued at 650 ◦ C and
possesses the highest {111}<112> SF along the plane parallel to the TB RT is caused by planar slipping on different {111} planes. The shearing
((111)[112]G1//(111)[121]T1), implying that an additional {111}< of γ′′ precipitates is clearly visible (Fig. 14b, c and e, f), indicating that
112> slip system could be activated for the twin and parent grain pair in the dislocation-precipitation shearing mechanism is predominant dur­
the HSA IN718 alloy under HCF loading at 650 ◦ C. ing cyclic deformation. The phenomenon of shearing of nanosized γ′′
To investigate the deformation behavior of precipitates during HCF precipitates also appears normally in tensile testing, even in the fatigue
loading, additional microstructure observations on the longitudinal process for IN718 superalloys fabricated by conventional methods
sections of the HCF fractures were performed, as illustrated in Fig. 13. [46–49].
Notably, under RT HCF loading, some carbides are fragmented and the
GB-δ particles basically maintain their features (Fig. 13a, b and d). 4.3. Fatigue life and limit
However, under HCF loading at 650 ◦ C, a small proportion of GB-δ
particles generated bending deformation (Fig. 13c and e), indicating a Conventionally, for IN718 alloys fabricated by traditional methods,
certain level of HT plastic deformability. Nevertheless, the influence of Nf decreases with increasing test temperature for the same σ a . The
the δ phase on σ w value reduction of the HSA IN718 alloy should be present study showed the same trends, i.e., the fatigue life of the HSA
small at both RT and 650 ◦ C because the σ w value of the HSA specimen is LDED IN718 alloy at RT was higher than that at 650 ◦ C for the same σ a
based on the results of the Basquin relationship. Moreover, the cumu­
lative failure at RT (bHSA− RT = − 0.2148) was lower than that at 650 ◦ C
(bHSA− 650◦ C = − 0.2049) for the HSA specimen. According to the Basquin
equation, σˊf ​ , which stands for fatigue damage tolerance, is another
factor influencing Nf . A larger σ ˊf ​ implies that more cycles can be
affected with the same failure rate. As shown in Table 3, both the
damage limit and σˊf ​ value of the HSA specimen at RT were higher than
those at 650 ◦ C. The σ ˊf ​ and b values at 650 ◦ C decreased by 42.6% and
4.6%, respectively compared to those at RT. Hence, the σ ˊf ​ value plays a
critical role in determining the fatigue life of the HSA specimen.
The HSA LDED IN718 specimen possesses a higher fatigue limit at RT
than at 650 ◦ C for the following reasons: (i) According to the empirical
rule, the σ w value is roughly estimated as (0.4–0.5)UTS for conventional
structural steels and alloys under fully reversed uniaxial loading
[50–52], although this empirical rule does not apply to
precipitation-hardened alloys (e.g., IN718 alloy). This may imply that
the high UTS of the HSA specimen at RT could lead to a higher σ w value
than that at 650 ◦ C, although there is a large error in predicting the σ w
value of the HSA LDED specimen. (ii) Distinct failure modes occur in
fatigued specimens between RT and 650 ◦ C because of the temperature
sensitivity of HCF. As shown in Figs. 8 and 9, the fatigue crack at RT was
mainly initiated from the surface slip bands. Nevertheless, the fatigue
crack at 650 ◦ C was mainly initiated in large crystallographic facets. (iii)
Environmental effects on the σw and Nf values of HSA LDED IN718
specimens should be considered, including oxygen embrittlement at the
GB [42,53] and specimen surface oxidation at HTs [28]. Hence, we
Fig. 12. EBSD map showing the HCF crack propagation path: (a, b) RT/500 speculate that failure modes, environmental effects, and the UTS influ­
MPa, Nf = 142,400 cycles; (c, d) 650 ◦ C/300 MPa, Nf = 638,800 cycles. ence the HCF behavior of HSA LDED IN718 alloys.

10
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Table 4
The {111}<110> slip system with the corresponding Schmid factors (SFs) for the twin and parent grain pair (Twin T1-Grain G1, in Fig. 12c).
{111}<110> slip system (111) (111) (111) (111)

[011] [101] [110] [011] [101] [110] [011] [101] [110] [011] [101] [110]

SF (Grain G1) 0.206 0.474 0.340 0.435 0.291 0.282 0.058 0.215 0.154 0.188 0.031 0.219
SF (Twin T1) 0.308 0.059 0.228 0.148 0.089 0.294 0.184 0.069 0.480 0.128 0.172 0.418

Table 5
The {111}<112> slip system with the corresponding Schmid factors (SFs) for the twin and parent grain pair (Twin T1-Grain G1, in Fig. 12c).
{111}<112> slip system (111) (111) (111) (111)

[121] [211] [112] [112] [211] [121] [121] [112] [211] [211] [121] [112]

SF (Grain G1) 0.419 0.130 0.283 0.332 0.272 0.152 0.099 0.495 0.171 0.249 0.263 0.157
SF (Twin T1) 0.352 0.248 0.398 0.168 0.084 0.298 0.498 0.198 0.225 0.401 0.233 0.278

Fig. 13. Microstructures of longitudinal sections of the HCF fractures: (a, b) RT/400 MPa, Nf = 1,069,500 cycles; (c) 650 ◦ C/320 MPa, Nf = 3,476,900 cycles; (d)
EDS spectra from the test position 1 in (b) showing the presence of (Nb,Ti)C particle; (e) EDS spectra from the test position 2 in (c) showing the presence of δ phase.

5. Conclusions (3) The CGs of the BG microstructure significantly reduced the HCF
properties of HSA specimens, thereby eclipsing the deleterious
This study investigated the microstructure, tensile, and HCF prop­ effect of the initial pore defects in the deposit. Under HCF loading
erties of the HSA LDED IN718 alloy at both RT and 650 ◦ C. The major at 650 ◦ C, the fatigue crack easily initiated and propagated along
conclusions are as follows: the TBs. Planar slip along the {111} planes with evidence of
sheared γ′′ phases occurred in HSA specimens tested in the HCF
(1) The grains of the HSA LDED IN718 alloy exhibited a BG distri­ regime at both RT and 650 ◦ C.
bution (10–500 μm) with submicron twins, and its second phase (4) The fracture surfaces of all the HCF specimens mainly consisted of
mainly contained abundant γ′′ and γ′ phases (~30 nm), dispersive the FCIZ, FCPZ, and FRZ. In the FCIZ, FCI was predominantly
and submicron carbides, and a few submicron and short rod-like initiated from the specimen surface at RT, while it was initiated
GB-δ phases. from the large crystallographic facets at 650 ◦ C. In the FCPZ, the
(2) The tensile properties of the HSA specimens at both RT and FCP of all specimens was mainly transgranular and presented the
650 ◦ C were superior to those of wrought IN718 (AMS 5662). The FS feature. In the FRZ, the failure mode of the tested specimens
σ w values of the HSA specimens at RT and 650 ◦ C were 385 and was microvoid coalescence ductile fracture.
280 MPa, respectively, which are much lower than those of
wrought IN718 at the same temperatures.

11
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

Fig. 14. Planar slip bands of IN718 alloy after HCF test revealing the sheared γ′′ particles: (a) bright-field TEM (BF-TEM) image adjacent to fracture of HSA specimen
fatigued to 1,069,500 cycles at RT/400 MPa, (b) high-resolution TEM (HR-TEM) image adjacent to slip band 1 (SB1) in (a), (c) Fast Fourier Transform (FFT) pattern
of γ′ ′ precipitates in (b); (d) BF-TEM image adjacent to fracture of HSA specimen fatigued to 3,476,900 cycles at 650 ◦ C/320 MPa, (e) HR-TEM image adjacent to slip
band 2 (SB2) in (d), (f) FFT pattern of γ′ ′ precipitates in (e).

CRediT authorship contribution statement [5] Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang,
W. Chen, T.T. Roehling, R.T. Ott, M.K. Santala, P.J. Depond, M.J. Matthews, A.
V. Hamza, T. Zhu, Additively manufactured hierarchical stainless steels with high
Xiaobin Yu: Conceptualization, Methodology, Investigation, Data strength and ductility, Nat. Mater. 17 (2018) 63–70.
curation, Writing – original draft. Xin Lin: Writing – review & editing, [6] B. Barkia, P. Aubry, P. Haghi-Ashtiani, T. Auger, L. Gosmain, F. Schuster,
Supervision, Funding acquisition. Zihong Wang: Investigation. Shuya H. Maskrot, On the origin of the high tensile strength and ductility of additively
manufactured 316L stainless steel: multiscale investigation, J. Mater. Sci. Technol.
Zhang: Formal analysis. Xuehao Gao: Formal analysis. Yufeng Zhang: 41 (2020) 209–218.
Validation. Yongming Ren: Writing – review & editing, Supervision. [7] Q.B. Nguyen, Z. Zhu, F.L. Ng, B.W. Chua, S.M.L. Nai, J. Wei, High mechanical
Weidong Huang: Supervision, Funding acquisition. strengths and ductility of stainless steel 304L fabricated using selective laser
melting, J. Mater. Sci. Technol. 35 (2019) 388–394.
[8] Z. Zhu, W. Li, Q.B. Nguyen, X. An, W. Lu, Z. Li, F.L. Ng, S.M.L. Nai, J. Wei,
Enhanced strength–ductility synergy and transformation-induced plasticity of the
Declaration of competing interest selective laser melting fabricated 304L stainless steel, Addit. Manuf. (2020)
101300.
[9] S. Sui, H. Tan, J. Chen, C. Zhong, Z. Li, W. Fan, A. Gasser, W. Huang, The influence
The authors declare that they have no known competing financial of Laves phases on the room temperature tensile properties of Inconel 718
interests or personal relationships that could have appeared to influence fabricated by powder feeding laser additive manufacturing, Acta Mater. 164 (2019)
the work reported in this paper. 413–427.
[10] F.C. Liu, H.M. Cheng, X.B. Yu, G. Yang, C.P. Huang, X. Lin, J. Chen, Control of
microstructure and mechanical properties of laser solid formed Inconel 718
Acknowledgements superalloy by electromagnetic stirring, Opt Laser. Technol. 99 (2018) 342–350.
[11] F.C. Liu, X. Lin, X.B. Yu, C.P. Huang, W.D. Huang, Evolution of interface and
crystal orientation of laser solid formed GH4169 superalloy during
This work was supported by the National Key Research and Devel­ recrystallization, Acta Metall. Sin. 50 (4) (2014) 463–470.
opment Program of China (Grant Nos. 2018YFB1105800, [12] X.B. Yu, X. Lin, F.C. Liu, L.L. Wang, Y. Tang, J.C. Li, S.Y. Zhang, W.D. Huang,
Influence of post-heat-treatment on the microstructure and fracture toughness
2018YFB1105900 and 2016YFB1100100); the National Natural Science
properties of Inconel 718 fabricated with laser directed energy deposition additive
Foundation of China (Grant No. 51323008). manufacturing, Mater. Sci. Eng., A 798 (2020) 140092.
[13] X.B. Yu, X. Lin, H. Tan, Y.L. Hu, S.Y. Zhang, F.C. Liu, H.O. Yang, W.D. Huang,
Microstructure and fatigue crack growth behavior of Inconel 718 superalloy
References manufactured by laser directed energy deposition, Int. J. Fatig. 143 (2021)
106005.
[1] T. DebRoy, H.L. Wei, J.S. Zuback, T. Mukherjee, J.W. Elmer, J.O. Milewski, A. [14] J.C. Lippold, S.D. Kiser, J.N. DuPont, Welding Metallurgy and Weldability of
M. Beese, A. Wilson-Heid, A. De, W. Zhang, Additive manufacturing of metallic Nickel-Base Alloys, Wiley, 2011.
components – process, structure and properties, Prog. Mater. Sci. 92 (2018) [15] T.M. Pollock, Alloy design for aircraft engines, Nat. Mater. 15 (2016) 809–815.
112–224. [16] A. Yadollahi, N. Shamsaei, Additive manufacturing of fatigue resistant materials:
[2] Y.M. Ren, X. Lin, X. Fu, H. Tan, J. Chen, W.D. Huang, Microstructure and challenges and opportunities, Int. J. Fatig. 98 (2017) 14–31.
deformation behavior of Ti-6Al-4V alloy by high-power laser solid forming, Acta [17] V. Prithivirajan, M.D. Sangid, The role of defects and critical pore size analysis in
Mater. 132 (2017) 82–95. the fatigue response of additively manufactured IN718 via crystal plasticity, Mater.
[3] Z. Zhao, J. Chen, H. Tan, G. Zhang, X. Lin, W. Huang, Achieving superior ductility Des. 150 (2018) 139–153.
for laser solid formed extra low interstitial Ti-6Al-4V titanium alloy through [18] S. Gribbin, S. Ghorbanpour, N.C. Ferreri, J. Bicknell, I. Tsukrov, M. Knezevic, Role
equiaxial alpha microstructure, Scripta Mater. 146 (2018) 187–191. of grain structure, grain boundaries, crystallographic texture, precipitates, and
[4] Z. Zhao, J. Chen, H. Tan, J.G. Tang, X. Lin, In situ tailoring microstructure in laser porosity on fatigue behavior of Inconel 718 at room and elevated temperatures,
solid formed titanium alloy for superior fatigue crack growth resistance, Scripta Mater. Char. 149 (2019) 184–197.
Mater. 174 (2020) 53–57.

12
X. Yu et al. Materials Science & Engineering A 825 (2021) 141865

[19] H.Y. Wan, Z.J. Zhou, C.P. Li, G.F. Chen, G.P. Zhang, Enhancing fatigue strength of [37] Editor Committee of Chinese Aerospace Material Manual, Chinese Aerospace
selective laser melting-fabricated Inconel 718 by tailoring heat treatment route, Material Manual: Deforming Superalloy and Casting Superalloy, second ed.,
Adv. Eng. Mater. 20 (10) (2018) 6. Standards Press of China, Beijing, 2001, p. 342.
[20] A.R. Balachandramurthi, J. Moverare, N. Dixit, D. Deng, R. Pederson, [38] N.E. Dowling, Mechanical Behavior of Materials, Engineering Methods for
Microstructural influence on fatigue crack propagation during high cycle fatigue Deformation, Fracture and Fatigue, fourth ed., Pearson, 2013.
testing of additively manufactured Alloy 718, Mater. Char. 149 (2019) 82–94. [39] G.R. Leverant, M. Gell, The influence of temperature and cyclic frequency on the
[21] S. Sui, J. Chen, E. Fan, H. Yang, X. Lin, W. Huang, The influence of Laves phases on fatigue fracture of cube oriented nickel-base superalloy single crystals, Metall.
the high-cycle fatigue behavior of laser additive manufactured Inconel 718, Mater. Trans. A 6 (1975) 367–371.
Sci. Eng., A 695 (2017) 6–13. [40] Y. Yamashita, T. Murakami, R. Mihara, M. Okada, Y. Murakami, Defect analysis
[22] D. Deng, R.L. Peng, H. Brodin, J. Moverare, Microstructure and mechanical and fatigue design basis for Ni-based superalloy 718 manufactured by selective
properties of Inconel 718 produced by selective laser melting: sample orientation laser melting, Int. J. Fatig. 117 (2018) 485–495.
dependence and effects of post heat treatments, Mater. Sci. Eng., A 713 (2018) [41] Y. Murakami, Metal Fatigue: Effects of Small Defects and Nonmetallic Inclusions,
294–306. Elsevier, Amsterdam, 2002.
[23] Y. Ono, T. Yuri, H. Sumiyoshi, E. Takeuchi, S. Matsuoka, T. Ogata, High-cycle [42] J.C. Stinville, E. Martin, M. Karadge, S. Ismonov, M. Soare, T. Hanlon,
fatigue properties at cryogenic temperatures in INCONEL 718 nickel-based S. Sundaram, M.P. Echlin, P.G. Callahan, W.C. Lenthe, V.M. Miller, J. Miao, A.
superalloy, Mater. Trans. 45 (2004) 342–345. E. Wessman, R. Finlay, A. Loghin, J. Marte, T.M. Pollock, Fatigue deformation in a
[24] Y.Y. Zhang, Z. Duan, H.J. Shi, Comparison of the very high cycle fatigue behaviors polycrystalline nickel base superalloy at intermediate and high temperature:
of INCONEL 718 with different loading frequencies, Sci. China Phys. Mech. Astron. competing failure modes, Acta Mater. 152 (2018) 16–33.
56 (2013) 617–623. [43] W.W. Kong, C. Yuan, B.N. Zhang, Investigations on cyclic deformation behaviors
[25] X.F. Ma, Z. Duan, H.J. Shi, R. Murai, E. Yanagisawa, Fatigue and fracture behavior and corresponding failure modes of a Ni-Based superalloy, Mater. Sci. Eng., A 791
of nickel-based superalloy Inconel 718 up to the very high cycle regime, J. Zhejiang (2020) 139775.
Univ. - Sci. 11 (2010) 727–737. [44] J. Miao, T.M. Pollock, J.W. Jones, Crystallographic fatigue crack initiation in
[26] K. Kobayashi, K. Yamaguchi, M. Hayakawa, M. Kimura, Grain size effect on high- nickel-based superalloy René 88DT at elevated temperature, Acta Mater. 57 (20)
temperature fatigue properties of alloy718, Mater. Lett. 59 (2) (2005) 383–386. (2009) 5964–5974.
[27] D. Fournier, A. Pineau, Low cycle fatigue behavior of Inconel 718 at 298K and [45] S.R. Yeratapally, M.G. Glavicic, M. Hardy, M.D. Sangid, Microstructure based
823K, Metall. Trans. A 8A (1977) 1095–1105. fatigue life prediction framework for polycrystalline nickel-base superalloys with
[28] N. Kawagoishi, Q. Chen, H. Nisitani, Fatigue strength of Inconel 718 at elevated emphasis on the role played by twin boundaries in crack initiation, Acta Mater. 107
temperatures, Fatig. Fract. Eng. Mater. Struct. 23 (2000) 209–216. (2016) 152–167.
[29] F.C. Liu, X. Lin, G.L. Yang, M.H. Song, W.D. Huang, Recrystallization and its [46] L. Xiao, D.L. Chen, M.C. Chaturvedi, Shearing of γ′ ′ precipitates and formation of
influence on microstructures and mechanical properties of laser solid formed nickel planar slip bands in Inconel 718 during cyclic deformation, Scripta Mater. 52
base superalloy Inconel 718, Rare Met. 30 (2011) 433–438. (2005) 603–607.
[30] L. Zhu, Z. Xu, Y. Gu, Effect of laser power on the microstructure and mechanical [47] P.M. Mignanelli, N.G. Jones, E. Pickering, O. Messe, C.M.F. Rae, M.C. Hardy, H.
properties of heat treated Inconel 718 superalloy by laser solid forming, J. Alloys A. Stone, Gamma-gamma prime-gamma double prime dual-superlattice
Compd. 746 (2018) 159–167. superalloys, Scripta Mater. 136 (2007) 136–140.
[31] X.S. Xie, C.M. Xu, G.L. Wang, J.X. Dong, W.D. Cao, R. Kennedy, in: E.A. Loria (Ed.), [48] H. Qin, Z. Bi, H. Yu, G. Feng, J. Du, J. Zhang, Influence of stress on γ′′ precipitation
Superalloys 718 625 706 Various Derivatives 2005, the Minerals, Metals & behavior in Inconel 718 during aging, J. Alloys Compd. 740 (2018) 997–1006.
Materials Society, 2005, pp. 193–202. [49] H.J. Lee, H.K. Kim, H.U. Hong, B.S. Lee, Influence of the focus offset on the defects,
[32] M. Sundararaman, P. Mukhopadhyay, S. Banerjee, Precipitation of the δ-Ni3Nb microstructure, and mechanical properties of an Inconel 718 superalloy fabricated
phase in two nickel base superalloys, Metall. Trans. 19A (1988) 453–465. by electron beam additive manufacturing, J. Alloys Compd. 781 (2019) 842–856.
[33] S. Sui, C.L. Zhong, J. Chen, A. Gasser, W.D. Huang, J.H. Schleifenbaum, Influence [50] S. Kevinsanny, O. Okazaki, Y. Takakuwa, Y. Ogawa, H. Funakoshi, S. Kawashima,
of solution heat treatment on microstructure and tensile properties of Inconel 718 H. Matsuoka, Matsunaga, Defect tolerance and hydrogen susceptibility of the
formed by high-deposition-rate laser metal deposition, J. Alloys Compd. 740 fatigue limit of an additively manufactured Ni-based superalloy 718, Int. J. Fatig.
(2018) 389–399. 139 (2020) 105740.
[34] K.S. Kim, T.H. Kang, M.E. Kassner, K.T. Son, K.A. Lee, High-temperature tensile [51] S. Nishijima, Statistical analysis of fatigue test data, J. Soc. Mater. Sci. Japan 29
and high cycle fatigue properties of inconel 625 alloy manufactured by laser (1980) 24–29.
powder bed fusion, Addit. Manuf. (2020) 101377. [52] Y. Tanaka, S. Okazaki, Y. Ogawa, M. Endo, H. Matsunaga, Fatigue limit of Ni-based
[35] J. Tan, X. Wu, E.H. Han, W. Ke, X. Liu, F. Meng, X. Xu, Role of TiN inclusion on superalloy 718 relative to the shear-mode crack-growth threshold: a quantitative
corrosion fatigue behavior of Alloy 690 steam generator tubes in borated and evaluation considering the influence of crack-opening and -closing stresses, Int. J.
lithiated high temperature water, Corrosion Sci. 88 (2014) 349–359. Fatig. 148 (2021) 106228.
[36] X.D. Xu, P. Liu, Z. Tang, A. Hirata, S.X. Song, T.G. Nieh, P.K. Liaw, C.T. Liu, M. [53] M. Reger, L. Remy, Fatigue oxydation interaction in IN 100 superalloy, Metall.
W. Chen, Transmission electron microscopy characterization of dislocation Transact. A 19A (1988) 2259–2268.
structure in a face-centered cubic high-entropy alloy Al0.1CoCrFeNi, Acta Mater.
144 (2018) 107–115.

13

You might also like