You are on page 1of 10

Engineering Fracture Mechanics 242 (2021) 107508

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


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

A machine-learning fatigue life prediction approach of additively


manufactured metals
Hongyixi Bao a, b, Shengchuan Wu b, c, *, Zhengkai Wu b, Guozheng Kang a, b, *,
Xin Peng b, Philip J. Withers c
a
Applied Mechanics and Structure Safety Key Laboratory of Sichuan Province, School of Mechanics and Engineering, Southwest Jiaotong University,
Chengdu 610031, China
b
State Key Laboratory of Traction Power, Southwest Jiaotong University, Chengdu 610031, China
c
Henry Royce Institute, Department of Materials, The University of Manchester, Manchester M13 9PL, UK

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

Keywords: The defects retained during laser powder bed fusion determine the poor fatigue performance and
Machine learning method pronounced lifetime scatter of the fabricated metallic components. In this work, a machine
Laser powder bed fusion learning method was adopted to explore the influence of defect location, size, and morphology on
Synchrotron X-ray computed tomography
the fatigue life of a selective laser melted Ti-6Al-4 V alloy. Both the high cycle fatigue post-
Fatigue life
Ti-6Al-4V alloy
mortem examination and synchrotron X-ray tomography were combined to acquire the geo­
metric features of the critical defects, which were trained using a support vector machine (SVM).
To accelerate the optimization process, the grid search approach with cross validation was
selected for fitting the model parameters. It is found that the coefficient of determination between
the predicted and experimental fatigue lives can reach up to 0.99, indicating that the SVM model
shows strong training ability.

1. Introduction

As a typical laser powder bed fusion method, selective laser melting (SLM) is capable of providing greater manufacturing efficiency
and flexibility [1-4]. On the other hand, light titanium alloys are used for weight-critical and stiffness-critical structures owing to their
superior strength-to-weight ratio and excellent corrosion resistance [3-5]. It has been found that the quasi-static service strength of
SLM processed Ti-6Al-4 V parts can meet the design criteria based on wrought materials. However, the manufacturing defects pro­
duced during the SLM typically lead to poor fatigue properties of near-net-shape components [6-8]. Thus, an in-depth understanding of
the effect of defects on the fatigue life of SLM processed parts is a prerequisite for use in both academic and industrial fields.
Representative manufacturing defects during additive manufacturing (AM) mainly include porosity and lack of fusion (LOF) defects
[4,6,8]. Gas pores are typically generated owing to gas entrapment or unstable keyhole and LOF defects from insufficient energy input
[9-11]. The LOF defects normally have irregular shapes and larger dimensions than the randomly distributed gas pores. Both porosity
and LOF defects have a detrimental effect on fatigue performance because they are stress concentrators promoting crack initiation.
Despite a low porosity (about 0.01%), the fatigue failure of SLM processed Ti-6Al-4 V samples predominantly originates from the
surface or near surface defects [8,12].

* Corresponding authors at: State Key Laboratory of Traction Power, Southwest Jiaotong University, Chengdu 610031, China (S.C. Wu), and
School of Mechanics and Engineering, Southwest Jiaotong University, Chengdu 610031, China (G.Z. Kang)
E-mail addresses: wusc@swjtu.edu.cn (S. Wu), guozhengkang@126.com (G. Kang).

https://doi.org/10.1016/j.engfracmech.2020.107508
Received 26 September 2020; Received in revised form 23 December 2020; Accepted 25 December 2020
Available online 31 December 2020
0013-7944/© 2020 Elsevier Ltd. All rights reserved.
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

Fig. 1. Microstructure (x-z sections) along the build direction for a sample manufactured with a laser power of 260 W and scanning speed of 1000
mm/s: (a) optical microscopy image and (b) inverse pole figure (IPF) electron backscattered diffraction (EBSD) map for the α-phase grains.

Previous studies have clearly indicated that defect location, size, and morphology have a detrimental influence on the fatigue life of
AM processed parts [13,14]. The combined effects of these geometric parameters determine the critical defects that trigger crack
initiation. The sample-to-sample variation in fatigue life is largely controlled by the position, size, and shape of the critical defects
[6,8,15]. To characterize the defect geometry, the conventional method is to examine post-mortem fracture surfaces using scanning
electron microscopy (SEM). With this technique, only two-dimensional (2D) defect features located on the specimen surfaces can be
obtained. Recently, high-resolution industrial and synchrotron radiation X-ray micro computed tomography (SR-μCT) technology have
been widely used to image three-dimensional (3D) defects and perform statistical analysis [16-19]. To effectively determine the
geometric features of the critical defects and to evaluate the critical defect-determined fatigue life, a very prospective approach is to
combine μCT with the machine learning (ML), which is a subset of artificial intelligence (AI) [20].
Machine learning techniques have demonstrated great potential for recognizing patterns in complex data from industrial com­
munities and recently for determining various mechanical and fatigue parameters of materials [21-23]. Algorithms such as artificial
neural networks (ANNs) and support vector machines (SVMs) have received a great deal of attention as promising ML approaches
[24,25]. They have greater computation accuracy and efficiency for small sample prediction and nonlinear regression analysis over
traditional statistical methods [26,27]. An increasing number of studies have successfully applied these original methods to fatigue life
prediction [25,28]. To date, very few studies have investigated the influence of AM defects on fatigue life using ML approaches.
This paper attempts to provide a fatigue life prediction method for SLM processed Ti-6Al-4 V parts using a combination of ML and
high-resolution SR-μCT. Prior to fatigue testing, SR-μCT was used to characterize the porosity and LOF defects, including their pop­
ulation, size, location, and morphology. After fatigue testing, post-mortem fractography and advanced SR-μCT were collaboratively
employed to acquire the geometric features of the critical defects. These critical defects were tentatively trained into the SVM model to
well correlate with the high cycle fatigue life. Finally, the defect-determined fatigue life was predicted followed by the lifetime
evaluation and validation.

2. Experimental procedure

2.1. The material

In this study, cylindrical Ti-6Al-4 V parts (Ø16 mm × 70 mm in height) were vertically built using a commercial EOS-M280
manufacturing system (build direction denoted as z). The Ti-6Al-4 V feeding powder with similar compositions had an average
diameter of 38 μm. Ref. [12] concluded the optimized processing parameters. After manufacturing, a post-SLM heat treatment at
500 ◦ C for 2 h followed by cooling under an argon atmosphere was adopted to substantially alleviate the residual stresses.
As shown in Fig. 1, the optical microstructure is characterized by large columnar β-phase grains and fine needle-like α-phase grains
with an average width of 1.5 μm. Furthermore, it is clearly observed that the α-lath width and colony size as well as the volume fraction
of the β phase determine the quasi-static mechanical properties of SLM processed Ti-6Al-4 V alloy [29,30], giving Young’s modulus of
105 GPa, ultimate tensile strength of 1267 MPa, 0.2% offset proof yield strength of 1094 MPa and engineering strain to failure of 7%
[12]. In the case of cyclic loading, the larger defects relative to the α grains are believed to be more important for crack initiation,
leading to poor fatigue properties and a pronounced scatter in fatigue life [13].

2.2. X-ray tomography

3D characterization of manufacturing defects within SLM processed Ti-6Al-4 V samples was conducted on the 13 W1 beam line at
the Shanghai Synchrotron Radiation Facility, China. Ten miniature dog-bone-shaped samples were placed 17 cm from a Hamamastu
Flash 4.0 sCMOS detector. The dimensions of these samples can be found in [8], and the SR-μCT imaged region was Ø2 mm × 2 mm in
height, which was the fracture region. The photon energy was 60 keV and the effective pixel size was 3.25 μm. A total of 720

2
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

Fig. 2. Typical SR-μCT reconstructed material volume, showing the extremely complex distribution inside SLM processed Ti-6Al-4 V specimens.

30
A2 +( A1 - A2 )
y=
1+( x / x0 ) p
25
A1 = 26.22
Relative frequency (%)

A2 = -0.94
20 x0 = 47.15
p = 5.74

15

10

0
0 10 20 30 40 50 60 70 80 90
Defect size area ( m)

Fig. 3. Statistical distribution of the dimensions of defects inside the as-received SLM test pieces, which indicates that almost all defects are less than
100 μm.

projections (rotation step size of 0.22◦ and exposure time of 3.5 s) were collected for volume reconstruction. Fig. 2 shows the
reconstructed volume of the samples. 3D defect reconstruction and statistical analysis were then performed using the commercial
software Avizo. The location, size, and morphology of the top twenty defects determined by the volume within each sample were
extracted.

2.3. Fatigue test

Axial fatigue tests were conducted on the above miniature dog-bone-shaped samples using an MTS Bionix 858 servo-hydraulic
testing machine after they were subjected to X-ray computed tomography. The sinusoidal loading was applied with a stress ratio of
0.1 at a frequency of 0.5 Hz. Fractographic analysis was performed using an SEM (FEI Quanta FEG 250) to identify the critical defects
within the failed specimens.

3. Results and discussion

3.1. Defect characterization

Both the porosity and the LOF defects have a detrimental effect on the fatigue performance because they can induce crack
nucleation, particularly under cycle fatigue loading. The level of fatigue degradation is largely controlled by the defect geometry,

3
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

18

16

14

Relative frequency (%)


12

10

0
0 100 200 300 400 500 600 700 800 900 1000
Distance to the surface ( m)

Fig. 4. The defect location statistical distribution by the distance to the specimen surface of as-received SLM test pieces, clearly showing almost all
the defects are located underneath the surface and few defects at the surface.

0.75

regular
Defect sphericity

0.60

0.45

inregular
0.30

0 15 30 45 60 75
Equivalent diameter ( m)

Fig. 5. Statistical distribution of the defect morphology inside the as-received SLM test pieces, which shows increasing irregularity with increasing
defect size.

including its location, size, and shape. The severity of their effect in order of increasing significance was morphology, dimension and
position [31]. Despite this, the combined effects of the three parameters determine the critical defects triggering crack initiation and
contribute to the pronounced scatter in fatigue life [32]. This section focuses on the statistical analysis of the manufacturing defects
using SR-μCT.
In the case of irregular AM defects in three dimensions, the well-recognized Murakami parameter, √area, has been widely used to
characterize the defects [33]. The indicator √area represents the square root of the projected area of the defect normal to the loading
direction. Fig. 3 shows the distribution of the projected defect areas (in the x-y plane). It is clearly observed that the defect population
decreases with increasing √area value. Around 85% of the defects have a √area between 10 and 50 µm. An exceedingly small portion
of the defects have a √area less than 10 µm, because the defects with a volume smaller than 100 µm3 are excluded to suppress the
image noise.
Previous studies have identified that the defect location is the most crucial factor controlling fatigue resistance [31,34]. It was
found that the largest defect would not initiate the dominant cracks even when it was relatively deep below the surface. This finding
has been reported for welded structures, casting alloys and AM processed metals [6,31,32]. It has been indicated that a surface defect a
diameter 10 times smaller than that of an interior defect is the crack site [33]. To characterize the defect location in the SLM formed
cylindrical test-pieces, the shortest distance from the defect centroid to the free surface, h, is adopted. As shown in Fig. 4, the
manufacturing defects are randomly distributed as expected.
To characterize the defect shape inside test pieces, the widely used morphological parameter of sphericity is employed [35]:

π 1/3 (6V)2/3
Sphericity = (1)
A

4
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

where V is the defect volume and A is the defect surface area.


The defect sphericity distribution for the as-received SLM test pieces is shown in Fig. 5. It is found that all the defects are far from
perfectly spherical defects with a sphericity of 1. Larger volume defects generally show a more irregular and tortuous shape. LOF
defects arising from insufficient energy input are typically larger and fairly complex in shape, leading them to be potential crack
initiators owing to higher stress concentrations. Smaller defects are rounder, which may in part be because they are gas pores or
because they are less well resolved.
The Murakami √area parameter, the shortest distance from the defect centroid to the free surface and the sphericity characterize
the size, location, and morphology of each defect. The data set can be established by combining the defect geometry with the fatigue
life of the samples. This data set becomes an important input for subsequent ML procedures.

3.2. Fatigue life prediction

An SVM is a supervised ML algorithm based on the structural risk minimization principle [36,37]. The SVM can map the inputs to
high-dimensional feature spaces using the kernel function, leading to high efficiency in nonlinear classification and regression analysis
[38]. This method is also called the support vector regression (SVR) when it is used for regression.
As an optimization algorithm, SVR typically defines a convex ε-insensitive loss function to be minimized and obtains the flattest
hyperplane with minimum error while balancing the model complexity [39,40]. Eq. (2) is the objective function, which can be
transformed into Eq. (3) according to Karush-Kuhn-Tucker conditions [41]:

1 ∑N
( )
min ‖ω‖2 + C ξn + ξ*n
2 n=1



⎪ yn − ωϕ(x) − b⩽ε + ξn







⎪ * (2)

⎨ − yn − ωϕ(x) − b⩽ε + ξn
s.t.



⎪ ξn , ξ*n ⩾0







⎩ n = 1, 2, ...N


N
[ ] 1∑ N
( )( ) ( )
min αn (ε − yn ) − α*n (ε + yn ) + αi − α*i αj − α*j K xi , xj
n=1
2 i,j=1
⎧ ∑n (3)

⎨ n=1
αn − α*n = 0
s.t.


αn , α*n ∈ [0, C]

where ω is the weight vector with the same dimension as ×, ϕ(x) is the mapping function of vector ×, b is the threshold, C is the penalty
parameter, and ε defines a margin of tolerance, in which no penalty is given to errors. ξn and ξ*n are relaxation factors, and an and a*n
are Lagrangian multipliers [41], which can be determined by:
N (
∑ )
ω= αi − α*j ϕ(xn ) (4)
i=1

This function in Eq. (2) can be expressed as:



N
( ) ( ) ∑
N
( ) ( )
f (x) = αi − α*i ϕ(xi )ϕ xj + b = αi − α*i K xi , xj + b (5)
i=1 i=1

where K (xi, xj) is the kernel function that satisfies Mercer’s condition [42]. The kernel function is one of the most critical factors in
determining the capability of the model. The radial basis function is selected owing to its high efficiency and accuracy compared to the
linear kernel and polynomial kernel models [43]:
( ) ( ⃦ ⃦2 )
K xi , xj = exp − g⃦xi − xj ⃦ (6)

where g is another critical parameter for regression analysis.


The three parameters ε, C, and g are intercorrelated and combine to determine the complexity and generalization ability of the SVR
model. Therefore, the determination of these parameters plays a significant role in the accuracy of the SVR model. Here a grid search
with cross-validation is adopted. Grid search is a widely used parameter tuning method, which is able to find the best combination of
parameters for the given model and test dataset by looping through different values of the parameters. Cross-validation is used to

5
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

Fig. 6. Overview of typical process selection and model evaluation using grid search with cross-validation to optimize the model parameters.

Table 1
Parameters determined by grid search with cross validation and the corresponding MSE.
c g ε MSE

32 2.5772 × 10-3 4.3819 × 10-6 0.5974

Fig. 7. Overview of the fatigue life prediction procedure based on the SVR model.

evaluate of the parameters from the grid search to determine which ones are superior. The datasets are separated into a test and a
training dataset. The performance of the SVR model can be evaluated by validating the training results with the test data.
K-fold (splitting into k-folds) cross-validation is a cross-validation technique in which each fold is used for testing, and the rest is
subjected to training. The average performance is obtained after k times validation. K-fold cross-validation has great potential for the
small dataset because it is able to maximize both the test and training data. Here k is set to 5 to pursue a reliable validation. An
overview of the grid search method is shown in Fig. 6. Mean square error (MSE) is used as a criterion for judging the quality of the
parameters:
N ( )2
1 ∑
MSE = ̂y n − yn (7)
N n=1

where ŷn is the value predicted using the SVR model, and yn is the experimental result. The MSE of the SVR model is believed to
minimized by the optimal combination of c, g, and ε, which are shown in Table 1.
Eight out of the ten data sets are randomly selected as the training set to obtain the trained SVR model. Note that the value of fatigue
life is standardized to z-scores before training. The remaining data is regarded as the testing set to provide an unbiased evaluation. An
overview of fatigue life prediction based on SVR is illustrated in Fig. 7.

3.3. Evaluation and validation

A comparison between the predicted and experimental high cycle fatigue life is shown in Fig. 8(a). It is noteworthy that the value of

6
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

(a) Experimental life


5 Predicted life
1.2x10

Fatigue life (cycles)


9.0x104

6.0x104

3.0x104

0.0
0 2 4 6 8 10
Sample Number
(b) Experimental life
Location & size
1.2x105 Location & morphology
Size & morphology
Fatigue life (cycles)

9.0x104

6.0x104

3.0x104

0.0
0 2 4 6 8 10
Sample Number

Fig. 8. Comparison between the experimental results and (a) the predicted fatigue life based on the SVR model containing all three defects features
and (b) the predicted fatigue life based on SVR model containing only two of the three defects features.

Table 2
Experimental fatigue life and that predicted by the SVR model along with the MSE and R2.
Sample number Experimental life [cycles] Predicted life [cycles] Difference [cycles] Difference [%] MSE R2

1 130,128 132,176 2048 1.57 7.0665 × 10-5 0.99418


2 15,013 15,497 484 3.22
3 102,336 103,979 1643 1.61
4 91,066 92,585 1519 1.67
5 20,614 21,137 523 2.54
6 48,305 48,417 112 0.23
7 81,765 81,639 126 0.15
8 35,147 35,245 98 0.28
9 (test set) 97,752 97,669 83 0.09
10 (test set) 70,809 70,225 584 0.82

fatigue life is reverted to the actual value, while the MSE is calculated before the reversion. It is clear from the comparisons that the
predicted fatigue life is in very good agreement with the experimental results, indicating that the SVR model has very high accuracy in
the regression analysis of SLM processed Ti-6Al-4 V samples. Although the fitting precision of the test set is lower than that of the
training set, the accuracy remains high.
Fig. 8(b) shows the predicted fatigue life from SVMs containing only two of the three defect features: location, size and morphology.
Through simple analysis, it is found that the more the curve with two features deviated from the original curve, the larger the lacking
feature the fatigue life. Therefore, the following conclusions can be drawn from Fig. 8(b): (1) The location, size, and shape of defects in
the Ti-6Al-4 V samples all influence the fatigue life; (2) Compared with size and morphology, location has a considerable larger effect
on the fatigue life of the samples, and the severity of their effect in order of increasing significance is morphology, dimension, and
position; (3) The consistency of this result with other research shows that the SVR model offers a strong anti-interference ability to
distinguish the severity of the influence of each feature, which is appropriate for fatigue life analysis. To quantify the overall per­
formance of the regression model, the coefficient of determination, R2, can be introduced:

7
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

(a)
Experimental life
5
Predicted life by SVR
1.24x10 Predicted life by KNN

Fatigue life (cycles)


9.30x10 4

6.20x10 4

3.10x10 4

0.00
0 1 2 3 4 5 6 7 8 9 10 11
Sample Number

(b)

105
Predicted life (cycles)

Predicted life by SVM


Predicted life by KNN
Equal life line
1.5 times dispersion
104
104 105
Experimental life (cycles)

Fig. 9. Comparison between the fatigue life predicted by the KNN and the experimental results.

Table 3
Experimental fatigue life and that predicted by the KNN model along with the MSE and R2.
Sample number Experimental life [cycles] Predicted life [cycles] Difference [cycles] Difference [%] MSE R2
-3
1 130,128 130,369 241 0.19 1.2736 × 10 0.96761
2 15,013 14,817 196 1.31
3 112,336 112,547 211 0.19
4 91,066 90,768 298 0.33
5 20,614 20,809 195 0.95
6 48,305 48,583 278 0.58
7 81,765 81,464 301 0.37
8 35,147 34,940 207 0.59
9 (test set) 97,752 97,559 193 0.20
10 (test set) 70,809 69,454 1355 1.91

( )2
∑N
n=1 ̂y n − y n
R2 = 1 − ∑N 2
(8)
n=1 n − y)
(y

where y is the mean value of the predicted results. R2 corresponds to 1 for a perfect fit. The experimental and predicted fatigue lives
with the corresponding MSE and R2 are shown in Table 2.
To verify the superiority of the SVR model, another widely used ML algorithm, the k-nearest neighbor algorithm (KNN), s used to
predict the fatigue lives of SLM processed Ti-6Al-4 V samples for comparison. As a commonly used non-parametric methods for
classification and regression, the KNN is known for its flexibility because it does not make any assumptions about the underlying data
[44,45]. The parameter k is determined by a grid search with cross-validation, which is the same as the parameter tuning in the SVR
model. A comparison between the CPU time for training and that for prediction is not performed because this study aims to predict
fatigue life prediction, which most commonly involves a small sample. A comparison between the predicted fatigue life using the KNN

8
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

model and the experimental results is shown in Fig. 9 and Table 3, respectively.
It is found that both the SVR and KNN models show the high accuracy for the prediction of the training set, and all of the samples
fall within a 1.5 times dispersion line. Furthermore, the SVR model shows the superiority in the prediction of the test set in comparison
with the KNN model. The generalization ability of the SVR model is strong in the defect-determined fatigue life regression analysis,
which is critical in the industrial community.

4. Conclusions

The fatigue life prediction of SLM processed Ti-6Al-4 V samples was performed based on machine learning and high resolution
synchrotron X-ray micro tomography. Manufacturing defects were primarily characterized in terms of their size, location, and
morphology. High cycle fatigue testing was conducted on these samples to obtain the critical defect-determined fatigue life. The
support vector regression model was used to predict the fatigue life with parameters extracted from a grid search with cross validation.
Based on this study, four conclusions can be drawn:

(1) The fatigue life is largely controlled by the geometric parameters of the critical defects. The relationships between the fatigue
life and the size, location, and morphology of the defects have non-linear characteristics and dispersion.
(2) The optimal parameters in the SVM model are acquired from a grid search using a cross validation method. The coefficient of
determination is 0.99418, indicating that the combination of model parameters is reasonable.
(3) The location, size, and shape of defects in the Ti-6Al-4 V samples all influence the fatigue life. The severity of their effect in order
of increasing significance is morphology, dimension, and position.
(4) Compared with the KNN model, the SVM model has a higher accuracy in the prediction of the test set in comparison. Moreover,
it is shown that the SVM model has a strong generalization ability in the prediction of the fatigue life of SLM processed Ti-6Al-4
V. The coefficient of determination between the predicted and experimental results reaches up to 0.99.

Thus the results show that the defect-determined fatigue life of SLM processed Ti-6Al-4 V samples is well predicted with relatively
high accuracy using machine learning, particularly by integrating the location, dimension and morphology of internal defects
extracted from advanced X-ray tomography. In ongoing works, other factors such as the residual stress and surface roughness, should
be considered to provide a more practical and reliable reference for the structural integrity assessment of AM formed metals.
Furthermore, more test samples and complex specimen geometries as well as collaborated residual stress, surface roughness, and
microstructure, should be employed to improve the SVM model.

Declaration of Competing Interest

The authors declared that there was no conflict of interest.

Acknowledgements

The authors sincerely appreciate the Joint Fund of Large-scale Scientific Facility of the National Natural Science Foundation of
China (U2032121). PJW acknowledges the European Research Council grant (CORREL-CT, 695638) and EPSRC grant EP/P006566/1.
SC Wu as an honorary professor is grateful for his visiting position at the Henry Royce Institute for Advanced Materials funded by
EPSRC through EP/R00661X.

References

[1] DebRoy T, Mukherjee T, Milewski JO, Elmer JW, Ribic B, Blecher JJ, et al. Scientific, technological and economic issues in metal printing and their solutions.
Nat Mater 2019;18(10):1026–32.
[2] Herzog D, Seyda V, Wycisk E, Emmelmann C. Additive manufacturing of metals. Acta Mater 2016;117:371–92.
[3] Liu S, Shin YC. Additive manufacturing of Ti6Al4V alloy: A review. Mater Des 2019;164:107552.
[4] Fatemi A, Molaei R, Simsiriwong J, Sanaei N, Pegues J, Torries B, et al. Fatigue behaviour of additive manufactured materials: An overview of some recent
experimental studies on Ti-6Al-4V considering various processing and loading direction effects. Fatigue Fract Engng Mater Struct 2019;42(5):991–1009.
[5] Kakiuchi T, Kawaguchi R, Nakajima M, Hojo M, Fujimoto K, Uematsu Y. Prediction of fatigue limit in additively manufactured Ti-6Al-4V alloy at elevated
temperature. Int J Fatigue 2019;126:55–61.
[6] Tammas-Williams S, Withers PJ, Todd I, Prangnell PB. The influence of porosity on fatigue crack initiation in additively manufactured titanium components. Sci
Rep 2017;7(1):7308.
[7] Edwards P, Ramulu M. Fatigue performance evaluation of selective laser melted Ti-6Al-4V. Mater Sci Engng, A 2014;598:327–37.
[8] Hu YN, Wu SC, Withers PJ, Zhang J, Bao HYX, Fu YN, et al. The effect of manufacturing defects on the fatigue life of selective laser melted Ti-6Al-4V structures.
Mater Des 2020;192:108708.
[9] Oliveira JP, Santos TG, Miranda RM. Revisiting fundamental welding concepts to improve additive manufacturing: From theory to practice. Prog Mater Sci
2020;107:100590.
[10] Wu SC, Yu C, Yu PS, Buffière JY, Helfen L, Fu YN. Corner fatigue cracking behavior of hybrid laser AA7020 welds by synchrotron X-ray computed
microtomography. Mater Sci Engng, A 2016;651:604–14.
[11] Bayat M, Thanki A, Mohanty S, Witvouw A, Yang SF, Thorborg J, et al. Keyhole-induced porosities in laser-based Powder Bed Fusion (L-PBF) of Ti6Al4V: High-
fidelity modelling and experimental validation. Addit Manuf 2019;30:100835.
[12] Hu YN, Wu SC, Wu ZK, Zhong XL, Ahmed S, Karakal S, et al. A new approach to correlate the defect population with the fatigue life of selective laser melted Ti-
6Al-4V alloy. Int J Fatigue 2020;136:105584.

9
H. Bao et al. Engineering Fracture Mechanics 242 (2021) 107508

[13] Biswal R, Zhang X, Shamir M, Mamun AA, Awd M, Walther F, et al. Interrupted fatigue testing with periodic tomography to monitor porosity defects in wire +
arc additive manufactured Ti-6Al-4V. Addit Manuf 2019;28:517–27.
[14] Barba D, Alabort C, Tang YT, Viscasillas MJ, Reed RC, Alabort E. On the size and orientation effect in additive manufactured Ti-6Al-4V. Mater Des 2020;186:
108235.
[15] Ferro P, Fabrizi A, Berto F, Savio G, Meneghello R, Rosso S. Defects as a root cause of fatigue weakening of additively manufactured AlSi10Mg components.
Theor Appl Fract Mec 2020;108:102611.
[16] Siddique S, Imran M, Rauer M, Kaloudis M, Wycisk E, Emmelmann C, et al. Computed tomography for characterization of fatigue performance of selective laser
melted parts. Mater Des 2015;3:661–9.
[17] Chapman TP, Kareh KM, Knop M, Connolley T, Lee PD, Azeem MA, et al. Characterisation of short fatigue cracks in titanium alloy IMI 834 using X-ray
microtomography. Acta Mater 2015;99:49–62.
[18] Du Plessis A, Yadroitsava I, Le Roux SG, Yadroitsev I, Fieres J, Reinhart C, et al. Prediction of mechanical performance of Ti6Al4V cast alloy based on microCT-
based load simulation. J Alloy Compd 2017;724:267–74.
[19] Wu SC, Xiao TQ, Withers PJ. The imaging of failure in structural materials by synchrotron radiation X-ray microtomography. Engng Fract Mech 2017;182:
127–56.
[20] Jordan MI, Mitchell TM. Machine learning: Trends, perspectives, and prospects. Science 2015;349(6245):255.
[21] Ye S, Li B, Li QY, Zhao HP, Feng XQ. Deep neural network method for predicting the mechanical properties of composites. Appl Phys Lett 2019;115(16):161901.
[22] Naik DL, Kiran R. Identification and characterization of fracture in metals using machine learning based texture recognition algorithms. Engng Fract Mech 2019;
219:106618.
[23] Ma XR, He XF, Tu ZC. Prediction of fatigue–crack growth with neural network-based increment learning scheme. Engng Fract Mech 2020;242:107402.
[24] Liu H, Jiao Y. Application of Genetic Algorithm-Support Vector Machine (GA-SVM) for Damage Identification of Bridge. Int J Comput Intell Appl 2012;10(04):
383–97.
[25] Figueira Pujol JC, Andrade Pinto JM. A neural network approach to fatigue life prediction. Int J Fatigue 2011;33(3):313–22.
[26] Romano S, Brandao A, Gumpinger J, Gschweitl M, Beretta S. Qualification of AM parts: Extreme value statistics applied to tomographic measurements. Mater
Des 2017;131:32–48.
[27] Du Plessis A, Yadroitsava I, Yadroitsev I. Effects of defects on mechanical properties in metal additive manufacturing: A review focusing on X-ray tomography
insights. Mater Des 2020;187:108385.
[28] Zhang M, Sun C, Zhang X, Goh PC, Wei J, Hardacre D, et al. High cycle fatigue life prediction of laser additive manufactured stainless steel: A machine learning
approach. Int J Fatigue 2019;128:105194.
[29] Leuders S, Thӧne M, Riemer A, Niendorf T, Trӧster T, Richard HA, et al. On the mechanical behaviour of titanium alloy TiAl6V4 manufactured by selective laser
melting: Fatigue resistance and crack growth performance. Int J Fatigue 2013;48:300–7.
[30] Rafi HK, Karthik NV, Gong HJ, Starr TL, Stucker BE. Microstructures and mechanical properties of Ti6Al4V parts fabricated by selective laser melting and
electron beam melting. J Mater Engng Perform 2013;22:3872–83.
[31] Serrano-Munoz I, Buffière JY, Mokso R, Verdu C, Nadot Y. Location, location & size defects close to surfaces dominate fatigue crack initiation. Sci Rep 2017;7:
45239.
[32] Fedor F, Manfred H, Norbert H, Nikolai K. Probabilistic fatigue-life assessment model for laser-welded Ti-6Al-4V butt joints in the high-cycle fatigue regime. Int
J Fatigue 2018;116:22–35.
[33] Murakami Y. Metal fatigue: Effects of Small Defects and Nonmetallic Inclusions. Oxford: Elsevier; 2002.
[34] Benedetti M, Fontanari V, Bandini M, Zanini F, Carmignato S. Low- and high-cycle fatigue resistance of Ti-6Al-4V ELI additively manufactured via selective laser
melting Mean stress and defect sensitivity. Int J Fatigue 2018;107:96–109.
[35] Kabir M, Richter H. Modeling of Processing-Induced Pore Morphology in an Additively-Manufactured Ti-6Al-4V Alloy. Materials 2017;10(2):145.
[36] Noble WS. What is a support vector machine? Nat Biotechnol 2006;24(12):1565–7.
[37] Nisbet R, Miner G, Yale K. Handbook of Statistical Analysis and Data Mining Applications. Boston: Academic Press; 2009.
[38] Sánchez AVD. Advanced support vector machines and kernel methods. Neurocomputing 2003;55(1–2):5–20.
[39] Awad M, Khanna R. Efficient Learning Machines. Berkeley: Apress; 2009.
[40] Xue D, Balachandran PV, Hogden J, Theiler J, Xue D, Lookman T. Accelerated search for materials with targeted properties by adaptive design. Nat Commun
2016;7(1):11241.
[41] French M. Fundamentals of Optimization: Methods, Minimum Principles, and Applications for Making Things Better. Cham: Springer; 2018.
[42] Suykens JAK, Vandewalle J, De Moor B. Optimal control by least squares support vector machines. Neural Netw 2001;14(1):23–35.
[43] Prajapati GL, Patle A. On performing classification using SVM with radial basis and polynomial kernel functions. 3rd Int Conf Emerg Trends Engng Technol
2010:512–5.
[44] Yao Z, Ruzzo WL. A Regression-based K nearest neighbor algorithm for gene function prediction from heterogeneous data. BMC Bioinform 2006;7(1):S11.
[45] Li Z, Zhang Q, Zhao X. Performance analysis of K-nearest neighbor, support vector machine, and artificial neural network classifiers for driver drowsiness
detection with different road geometries. Int J Distrib Sens Netw 2017;13(9):812126107.

10

You might also like