Data Driven Modeling of Mechanical Properties For 17 4 PH Stainless

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/359888808

Data-Driven Modeling of Mechanical Properties for 17-4 PH Stainless Steel


Built by Additive Manufacturing

Article in Integrating Materials and Manufacturing Innovation · April 2022


DOI: 10.1007/s40192-022-00261-8

CITATIONS READS

5 191

4 authors, including:

Michael Porro Akanksha Parmar


Purdue University Purdue University
1 PUBLICATION 5 CITATIONS 1 PUBLICATION 5 CITATIONS

SEE PROFILE SEE PROFILE

Yung Shin
Purdue University
386 PUBLICATIONS 18,843 CITATIONS

SEE PROFILE

All content following this page was uploaded by Akanksha Parmar on 26 July 2022.

The user has requested enhancement of the downloaded file.


Integrating Materials and Manufacturing Innovation (2022) 11:241–255
https://doi.org/10.1007/s40192-022-00261-8

TECHNICAL ARTICLE

Data‑Driven Modeling of Mechanical Properties for 17‑4 PH Stainless


Steel Built by Additive Manufacturing
Michael Porro1 · Bin Zhang1 · Akanksha Parmar1 · Yung C. Shin1

Received: 10 December 2021 / Accepted: 23 March 2022 / Published online: 11 April 2022
© The Minerals, Metals & Materials Society 2022

Abstract
This study examines the link between microstructure and mechanical properties of additively manufactured metal parts
by developing a predictive model that can estimate properties such as ultimate tensile strength, yield strength, and elonga-
tion at fracture based upon microstructural data for 17-4 PH stainless steel. The main benefit of the approach presented
is the generalizability, as necessary testing is further reduced in comparison with similar methods that generate full pro-
cess–structure–property linkages. Data were collected from the available literature on AM-built 17-4 PH stainless steel,
in-house tensile testing and imaging, and testing conducted by an AM company. After standardizing the image size and grain
boundary extraction via image processing, the features such as grain size distributions and aspect ratios were extracted. By
using artificial neural networks, relationships were established between grain size and shape features and corresponding
mechanical properties, and subsequently, properties were predicted for novel samples to which the network had not previ-
ously been exposed. The model produced correlation coefficients of R2 = 0.957 for ultimate tensile strength, R2 = 0.939 for
yield strength, and R2 = 0.931 for fracture elongation. These results demonstrate the efficacy of predictive models that focus
upon microstructure–property relationships and highlight an opportunity for further exploration as predictive modeling of
metal additive manufacturing continues to improve.

Keywords Additive manufacturing · Data-driven model · 17-4 PH stainless steel · Mechanical properties · Microstructure

Introduction Unfortunately, several hurdles impede the proliferation


of additive manufacturing. One such barrier is the high
The ability to additively manufacture metal parts presents cost associated with determining the requisite operating
an opportunity to spark a major change in a variety of parameters for printing metal materials by AM to achieve
industries. From producing aircraft parts to turbine blades the desired microstructural and mechanical properties. This
for energy production, additive manufacturing of metal has high upfront cost significantly hinders widespread adoption
the potential to be a transformative production process. Its in the industry. The objective of this study is to establish a
advantages include greatly shortened lead time for prototyp- procedure to reduce this cost by using data-driven modeling
ing, a reduction in material waste, the ability to build unique to predict the mechanical properties of AM-built 17-4 PH
and complex geometries, and a reduced need for assembly stainless steel parts. Without predictive modeling, process
upon build completion. Given the lack of need for tooling, parameters for optimal or even successful operation must
along with the ability to build parts in small quantities with- be determined experimentally or by using the parameters
out as many fixed manufacturing costs as traditional pro- already developed by those who produce AM systems,
cesses, metal additive manufacturing (AM) is an optimal which limits the potential for both the rapid implementation
choice in many cases for small volume production runs. of AM and customization of part performance. By utilizing
the method outlined in this study, the degree to which trial-
and-error strategies must be applied in the development of
* Yung C. Shin
shin@purdue.edu additively manufactured metal parts to meet required design
and performance specifications can be reduced. Ideally, with
1
Center for Laser‑Based Manufacturing, School sufficiently advanced predictive tools, parts of specific ten-
of Mechanical Engineering, Purdue University, sile strength or ductility, for example, could be produced
West Lafayette, IN 47907, USA

13
Vol.:(0123456789)
242 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

with minimal experimentation based on known relationships and processes used. Such data-driven modeling can also
between process parameters, microstructures, and mechani- utilize the widely available but scattered microstructure
cal properties. data with corresponding mechanical properties in the lit-
Traditional methods of determining mechanical proper- erature, thereby further reducing the expenses and costs of
ties include tensile or compression testing of many sam- generating requisite data. If a suitably trained data-driven
ples to determine mechanical characteristics such as yield model, such as an artificial neural network, is established
strength, ductility, and ultimate tensile/compressive strength, and exposed to an appropriate dataset, mechanical proper-
but this process is expensive and time-consuming in compar- ties can be predicted within a range based upon knowledge
ison with predictive methods. Given the wide range of addi- of microstructural characteristics without the need for much
tive manufacturing procedures that have recently become further testing.
available as well as the number of materials, performing Applications of machine learning models in the context
mechanical testing of all the parts can become a limiting of manufacturing are well documented in the literature. Neu-
factor for transitioning new AM methods from niche or ral networks have been applied to predicting the mechani-
experimental usage to more widely accepted industry usage. cal properties of traditionally manufactured steels since
In comparison with tensile testing of many individual ten- the early 2000s. For example, Tenner et al. [1] predicted
sile bars and experimentally determining every mechanical the effects of heat treatment on traditionally manufactured
property, a predictive approach to determine the mechanical steels, and Sterjovski et al. [2] similarly used backpropa-
properties based on known microstructures can provide a gation network models to predict a variety of mechanical
faster and cheaper solution. properties for heat-treated and traditionally manufactured
Although physics-based multi-scale modeling is a prom- steels. The studies concluded that given knowledge of com-
ising strategy, high-fidelity modeling approaches capable position, test conditions, and post-processing parameters, a
of accounting for microstructural details are presently very neural network could accurately predict resultant properties.
computationally expensive and cannot be applied to large- Since these early applications of neural networks to material
scale parts. This is because, in order to predict the macro- property prediction, the practice has proliferated to include a
mechanical behavior of materials, the physics-based meth- variety of strategies and outcomes (see Table 1).
ods may need to iteratively simulate the interactions among Techniques have been developed to simulate entire
microelements (i.e., molecules, grains). Therefore, even microstructures based on powder characteristics, deposi-
simulating for a small volume of material may take con- tion methods, and heating conditions, but these require a
siderable time. On the contrary, the data-driven approaches trade-off between model fidelity and computational expense
could learn the linkage between the overall characteristics [3]. Popova et al. [4] developed a strategy for obtaining pro-
of microstructures and mechanical properties, given that cess–structure–property linkages using simulated micro-
a sufficient and representative dataset for the material of structures generated by the Monte Carlo method. This
interest can be collected. Once established, the data-driven method specifies a particular set of operating conditions as
model can provide a quicker way of predicting mechanical an input for each model rather than generalizing its approach
properties based on given microstructure information for to a wide variety of AM processes applied to a given mate-
the same material regardless of the operating parameters rial. In 2020, Herriot and Spear [5] used convolutional

Table 1  Summary of common methods for predicting properties in additively manufactured materials
Strategy Data upon which predictions are made Form of prediction Applicable materials

Characteristics used in this study


Machine-learning models Microstructure of built part Yield strength Metals
Test data of built parts Ultimate tensile strength
Fracture strain
(properties as observed in tensile tests)
Characteristics used in other studies
Mechanistic models [3] Known material properties and composition Simulated microstructure [4, 5] Plastics [8]
information [1, 2]
Trial and error [1] Test conditions [1, 2] Fatigue behavior [6] Organic materials
Multiphysics modeling [4, 5] AM method and process parameters [3, 8] Property maps [4, 5, 7] Composites [7]
Governing equations [7, 11] Post-processing data [2] Young’s modulus [10] Ceramics
FEM simulations [7, 11] Thermal history [9]

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 243

neural networks and synthetic 3D microstructures (gener- produce PSP linkages. This study took a modular approach
ated using Dream3D) to predict microstructure-dependent to prediction and utilized several distinct data sets to arrive
physical properties of 316L steel built by simulated directed at each prediction of cycles to failure.
laser deposition (DLD). This study was able to predict yield While it has been well established that neural networks
strength with an R2 value of up to 0.95 using synthetic 3D can predict mechanical properties of steels and even account
images and grain orientation information. for the effects of post-processing, the application of these
Some challenges with these methods include the possible techniques to additively manufactured steels is compara-
variance in process conditions in different sections of an tively new and unexplored. A 2015 review published by Gao
AM part, as well as the need to account for defects in AM et al. [12] asserts that the imprecise nature of current meth-
metal parts, though many studies have been published on ods for determining optimal parameters for metal AM pro-
this particular subject. As an alternative to simulating micro- cesses is one of the main difficulties hindering the adoption
structures, basing predictions on observable microstructure of metal AM as a widespread procedure. It can be argued
characteristics allows for the overall computational cost of that simulations and machine learning-based approaches
producing a machine learning model to be lowered. One to process optimization must be further developed to make
such study conducted by Wan et al. [6] focused on fatigue metal AM a more viable manufacturing process in a broader
properties of AM metal parts and relied on data samples range of industries, despite difficulties associated with
collected near stress concentrations on a built part to acquire generating high-fidelity models. A particular strength of
data that could be used to predict future performance. Yang machine learning is that a variety of accessible programs
et al. [7] developed a method to predict mechanical proper- such as MATLAB exist, which can be used to train data-
ties as well as local features such as stress concentrations driven models. Thus, the procedure is replicable and can be
in 2D composites using generative adversarial networks adapted to a user’s specific needs. No matter what method
(GANs). Similar to the method outlined later in this study, is chosen; however, some amount of validation must be con-
this method used a collection of images as part of the input ducted by comparing predicted results with those observed
data for the neural network. This model was able to predict in laboratory testing.
stress field behavior for a coarse resolution. Additive manufacturing can be approached from a
Machine learning (ML) methods for tensile test result three-tiered perspective. The first tier consists of material
prediction have been applied with great success to other and feedstock properties, build conditions and process,
additively manufactured materials, such as plastics. In one and post-processing parameters. The second tier consists
such study, Bayratkar et al. [8] were able to achieve an R2 of microstructural characteristics which can be observed
value of over 0.999 using a similar network structure and after all printing and processing are complete, and the third
the same network algorithm as used in the current study. tier consists of the mechanical properties and performance
However, the study by Bayrathkar et al. based predictions of a complete part. This model focuses on the connection
on processing parameters rather than microstructure data. between tiers two and three, as opposed to specifying the
Data types to be used as input information for a neural net- model in terms of whole product–process–performance link-
work are a key consideration in applying ML to property ages. The advantage of this type of model is that once estab-
prediction, as the factors that can influence these traits are lished, it is applicable to the same material regardless of
numerous. processes and the process parameters used and hence serves
A variety of creative approaches have been combined as a universal model. By generalizing the microstructure and
with the capabilities of machine learning models to predict property relationship that can be applied to a given material,
or measure property data. Xie et al. [9] utilized thermal his- this model can be used to make satisfactory predictions for a
tory data obtained by IR scanning of the 3D-printed parts whole set of linkages instead of necessitating a new model
undergoing the cooling process to train convolutional neural for each individual process change for a given material. In
networks to predict UTS values (among other properties) at fact, a wide variety of microstructure types and processing
several points within the part body. This process has a great parameters would strengthen this sort of model, meaning
effect on a part’s eventual microstructure and represents one that as experimentation continues in the field, the accuracy
predictive method that can be generalized to apply to many of this model can improve.
different AM procedures. Lu et al. [10] developed a method This study focused on predicting various mechanical
of property measurement using instrumented indentation properties of 17-4 PH stainless steel, specifically, and did
and simulated test data to accurately measure properties of so based on both microstructural images of built samples and
metals, particularly Young’s modulus. One advantage of this data acquired from tensile testing of AM parts. The model
test method is that it can be nondestructive for large samples. employs machine learning in the form of artificial neural
Yan et al. [11] combined a variety of approaches, including networks as opposed to FEM simulations, derived govern-
FEM simulations and powder spreading characteristics to ing equations, or mechanistic models. Finally, this study

13
244 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

presents predictions of various attributes of mechanical images and corresponding mechanical properties for these
properties, rather than focusing only on one and the model samples. Ultimate tensile strength, yield strength, and frac-
output includes predicted values for ultimate tensile strength, ture strain data were collected. As demonstrated in Fig. 1,
yield strength, and fracture elongation. Other models as images used in the study were meant to show many grains in
listed in Table 1 have produced simulated microstructures order to accurately demonstrate generalizable microstructure
or predicted fatigue behavior; however, the current study characteristics for the whole sample. These microstructure
focuses on predicting resultant properties based upon images characteristics, such as grain size distribution, grain orien-
containing microstructure data regardless of processes and tation, and grain shape, are ultimately what determines the
process parameters. mechanical properties of the sample. For example, one such
relationship is the Hall–Petch effect, which shows the cor-
relation between grain size results and yield strength. Based
Experiments, Data Collection, on the microstructure information made available to the arti-
and Preprocessing ficial neural network, it iteratively derives similar relations,
which form the basis of predicted microstructural properties.
Data/Sample Collection Figure 1 shows an example of data collected from the
literature. Table 2 contains information regarding the col-
Some data used in this study were collected from the pre- lected dataset and sources utilized to collect it. The data
viously published literature in the form of microstructure collected from available literature consisted of tests con-
ducted on selective laser melting (SLM) and laser powder
bed fusion (LPBF).

Physical Sample Preparation

Physical samples were sourced from two companies: Digi-


tal Metal and Innovative 3D Manufacturing. Twenty-three
samples were tested in total, eight from Digital Metal and
15 from Innovative 3D Manufacturing. The samples from
Digital Metal were built using binder jetting and were built
in the flat orientation (see Fig. 2). The samples from Innova-
tive 3D Manufacturing were built using a spot-heating-based
AM process (Renishaw AM), and five samples were built
in the flat, upright, and vertical orientations, respectively
(see Fig. 2).
Two regions in the gage section of tested samples were
chosen to examine the microstructure morphology: one
closer to the gripping area and one closer to the frac-
Fig.1  An example of collected micrograph illustrating grain mor- ture as shown in Fig. 3. For microstructure characteri-
phology [13] zation, the tested samples were cross-sectioned with an

Table 2  Data from the related Sources UTS range (MPa) YS range (MPa) Ductility (%) Manufactur-
literature ing process
used

Yadollahi et al. [13] 940–960 635–725 3.33–4 SLM


Rafi et al. [14] 944 570 50 SLM
Yadollahi et al. [15] 940–1060 580–650 2.8–14.5 LPBF
Pasebani et al. [16] 990–1368 500–1255 2–5.5 SLM
Hsu et al. [17] 1003–1352 821–1229 7.3–13.8 SLM
Mahmoudi et al. [18] 950–1010 600–650 4–6.4 SLM
Alnajjar et. al. [19] 930–1130 850–1050 20.8–25.9 SLM
Yadollahi et al. [20] 1410 1250 11 SLM
Wang et al. [21] 744–882 541–568 30–42 SLM

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 245

Fig. 4  Image of etched sample from Digital Metal

Fig. 2  Sample build orientations


Tensile Test Experiments
­Al2O3 abrasive cutting wheel at 3000 RPM and mounted
in bakelite. The mounted samples were ground using The samples in this study were prepared using the ASTM
#180–#1200 SiC papers followed by mechanical polish- Standard E8/E8M-16aϵ1 for mechanical tensile testing. Sub-
ing with 3 μm diamond paste and 0.5 μm colloidal silica. size specimens were produced at a thickness of 0.04–0.0595
The samples were then etched with Kalling’s reagent no. inches, below the maximum thickness of 0.25 inches for
1 (ASTM 095) by swabbing for 30 s to reveal the micro- these samples. For five samples from Innovative 3D Manu-
structure. Kalling’s reagent no. 1 darkens martensite, facturing, these supports were attached to the gage section
slightly colors ferrite, and does not attack austenite. of the specimens. Printing support structures that could not
The microstructure was observed using an optical be removed by hand were removed using a file. Three parts
microscope (Nikon Eclipse LV150) and a scanning elec- had these removed (before measurement and testing) by fil-
tron microscope (JEOL-6400). All images on the opti- ing, and two samples were tested with these supports still
cal microscope were taken at 50X magnification, and all attached, but they were found to not affect testing.
images in the SEM were taken at 500X magnification and The samples were tested using a United Test Systems
25 kV power. 10kN Universal Testing Machine according to ASTM
Images collected using the optical microscope are Standard E8/E8M-16aϵ1. A strain rate of 12.7 mm/min (0.5
shown in Figs. 4 and 5. These two images represent inches/min) was used. An extensometer was used to track
samples from Digital Metal and Innovative 3D Manu- the first 3% of engineering strain, and the rest was calcu-
facturing, respectively. In addition, 55 images shared by lated using relations of total grip displacement and cross-
Digital Metal were obtained similarly as a result of the sectional area of gage and grip sections. The engineering
company’s in-house testing, and a sample image from this stress–strain curve of each test sample was obtained from the
dataset is shown in Fig. 6. output of the tensile testing system and then converted to the
true stress–strain curve to evaluate the sample’s mechanical
properties. The fitted linear line of elastic deformation was

Fig. 3  Spots for microstructure


characterization in the tensile
specimens

13
246 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

Fig. 5  Image of etched sample from Innovative 3D Manufacturing Fig. 7  Mechanical property extraction from the stress–strain curve

Data‑Driven Modeling of Mechanical


Properties

Grain Morphological Feature Extraction

Processing of collected images was carried out using the


image processing software ImageJ with Fiji and MorphoLibJ
add-ons installed. All initial processing consisted of two
main steps: obtaining a scale factor to convert from pixels
to micrometers and obtaining watershed lines representative
of grain boundaries in the microstructure images. Processing
parameters vary greatly between images, and this process
must be iterated to achieve accurate grain boundary images.
As the goal of data processing is to obtain a histogram
of grain sizes, it is crucial to be able to convert all images
to the same scale. To obtain the scale factor, the straight-
Fig. 6  Image of etched sample provided by Digital Metal line tool was used along with the “Measure” function. The
scale factor for laboratory-tested samples was 683 pixels
to 1000 µm, but the range of scales in this experiment var-
shifted to the right by a strain of 0.002 (0.2%), and its inter- ied from 220 pixels to 10 µm in some images from Digital
section with the true stress–strain curve was recognized as Metal to 50 pixels to 100 µm in some images collected from
the yield point, from which the yield strength was extracted. the literature. Generally, higher-resolution images improve
In addition, the ultimate tensile strength and fracture strain the software’s ability to accurately detect grains and draw
were extracted as the maximum values of the true stress and boundaries.
strain, respectively. The fracture strain will be used as the Watershed lines had to be generated to identify grain
measure of ductility in this work. An example of mechanical boundaries in a manner that can be fed into a grain size his-
property extraction is shown in Fig. 7. togram analysis. There are two methods applied to segment
images. The first step, which is applied to all image types, is
to convert to a grayscale 32-bit image format. Additionally,
the scale bars embedded in images were removed to elimi-
nate noise from grain boundary detection.
For non-EBSD images, the “Find Maxima” tool was used
for segmentation. This tool is used to draw boundaries about
specific regions whose brightness values differ from their
surroundings significantly enough to exceed allowable noise

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 247

tolerance. When utilized with optical microscope images systematically, the histogram of grain diameter is extracted
of material microstructures, a grain boundary map could as a set of morphological features, as opposed to only cal-
be developed. Iteration of this process is required to tune culating the mean grain diameter. The empirical Hall–Petch
parameters such as image contrast and noise tolerance to constants of mild steel [22] (σ0 = 70 MPa, ky = 0.74 MPa/
obtain the most accurate segmentation possible. m0.5) were adopted to sketch the Hall–Petch curve. Accord-
For EBSD images, a different tool was used from the ing to the Hall–Petch effect, the yield strength is more sen-
MorphoLibJ plug-in. This plug-in’s Morphological Segmen- sitive to grain size and thus the Hall–Petch curve is steeper
tation menu can be used to preview boundaries and execute when the grain diameter is small, while the strength will
segmentation. This is more effective when applied to EBSD be nearly constant and the curve will be flat when the grain
images because the Morphological Segmentation tool can diameter is large. This nonlinear effect of grain size could be
detect boundaries as brightness minima, and for this image better addressed by converting both yield strength and grain
format, grain boundaries will appear darker than their sur- diameter to log scales. As shown in Fig. 8, on the Hall–Petch
roundings. Again, the process is iterative, and parameters curve in log scale (i.e., log(σy − σ0) vs. log(d)), the yield
related to noise and contrast were adjusted such that an accu- strength log(σy − σ0) decreases linearly as the grain diameter
rate map of the grain structure is achieved. log(d) increases. Therefore, if the grain diameter bins in the
The morphological characteristics (e.g., size, shape, and histogram are chosen uniformly in log scale, the expected
orientation) of the grains formed during the AM process reduction in yield strength due to the Hall–Petch effect will
have a fundamental impact on the mechanical properties be the same in each bin. This strategy was adopted to select
of fabricated parts. The Hall–Petch relationship states that the histogram bins.
under the grain boundary strengthening effect, the material’s To extract the grain size histogram, the area of each grain
yield strength reaches the maximum value at a certain grain and its equivalent diameter
√ by assuming the grain to be per-
size (typically 10 nm). Then, as the grain size continues to fectly circular (i.e., 4∗Area
) was computed from the
increase, the grain boundaries start to slide, and hence the
𝜋
extracted grain boundaries. Throughout the whole 17-4 PH
yield strength drops. The inverse relationship between yield steel dataset from both the literature and in-house testing,
strength and the average grain size can be described by the the minimum grain diameter was found to be 0.37 µm and
Hall–Petch equation [22]: most grains were below 100 µm with only a few above
ky 100 µm. Therefore, a set of uniform histogram bins ranging
𝜎y = 𝜎0 + √ from − 0.5 (i.e., 0.316 µm) to 2 (i.e., 100 µm) with 0.25
d width in log scale were adopted to discretize the grain size.
In addition to these 10 uniform-width bins, a bin to account
where σy is the yield stress, σ0 and ky are constants specific
for all the grains above 100 µm was also included. Then, the
to each material, and d is the average grain diameter. How-
probability of a grain falling into each bin was evaluated by
ever, only considering the average grain diameter may not
counting the frequencies of grain diameters to obtain the
fully characterize the grain size effect and the probability
histogram. For a microstructure image with recognized grain
distribution of grain diameter should be more informative
boundaries, the probability of each histogram bin forms a
[23]. In this work, to analyze the effect of grain size more

Fig. 8  Hall–Petch curve in log


scale and the selected grain size
bins

13
248 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

feature vector to characterize the Hall–Petch effect of grain alter these two ratios. In Fig. 10, an elongated grain and a
size. In Fig. 9, the grain size histogram by counting the near-circular grain are compared. As can be seen, the two
grains in the whole dataset is presented. circularity features exhibited distinct values for these two
In addition to the grain size features from the histogram, grains and thus provide grain shape information comple-
two grain shape features used to measure the circularity of mentary to the grain size histogram. The correlation between
grains were also evaluated, which were the aspect ratio (the these circularity features and the variation of mechanical
ratio of major axis length to minor axis length, measured properties could be learned by the data-driven model. In
from the ellipse that has the same normalized second central total, there will be 13 grain morphological features extracted
moments as the grain region) and the ratio of grain perimeter for each microstructure image: the probabilities of a grain
to diameter. For a perfect circular grain, these two ratios falling into each of the 11 grain diameter histogram bins, the
should be 1 and 𝜋 , respectively. The motivation of using mean aspect ratio averaged over all the grains on the image,
these two features is that the presence of phases with differ- and the mean perimeter-to-diameter ratio.
ent grain shapes, such as the martensite phase with elongated
grains, will affect the mechanical properties and meanwhile Neural Network for Mechanical Property Prediction

Building a data-driven model to predict a target quantity


(mechanical property) based on a set of predictor features
(grain size, aspect ratio, etc.) is fundamentally a regres-
sion problem and can be done using a variety of supervised
machine learning models. There are a variety of ML mod-
els for regression problems, like an artificial neural net-
work, random forest, and support vector regression. With a
properly designed model structure, all these models could
achieve a good performance. Among these models, the arti-
ficial neural network is adopted in this work, owing to its
well-developed training technique, customizable model size,
and stable performance on various problems. Neural net-
works enable the learning of nonlinear relationships between
grain morphology and mechanical properties, which may
be actually more complex than the simple guideline of the
Hall–Petch equation and hence could not be well addressed
Fig. 9  Example histogram of the whole dataset by simple models like linear regression.

Fig. 10  Illustration of the circu-


larity features

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 249

Considering that the number of data samples obtained fell within a certain size range. Also included were features
from the literature and tensile tests was limited, the data- to describe grain shape (aspect ratio and perimeter-to-diam-
set was augmented by dividing each grain boundary image eter ratio). The predicted properties considered in this work
extracted in “Grain morphological feature extraction” sec- were ultimate tensile strength (UTS), yield strength (YS),
tion into a series of sub-images. A moving window with and fracture strain. The statistics of the input and output
50% overlapping in each moving step was used to crop sub- variables in the collected dataset are summarized in Table 3.
images from each full microstructure image. Each sub-image Since the data used for neural network modeling in this work
represented a sub-domain of the original sample and was was collected from a wide range of literature and the in-
assigned with the same mechanical properties as the original house testing of AM samples using different processes and
one. Grain morphological features were extracted for each built in different directions, it is believed that the dataset is
sub-image, i.e., when cropping images, if the centroid of a representative of the additively manufactured 17-4 PH steels
recognized grain is located in the current cropping frame, in practice.
this grain will be counted in the grain size histogram of this In this work, the Levenberg–Marquardt algorithm was
sub-image. This data augmentation process increased the used to develop training functions. This algorithm is strong
total number of data samples to about 5800, and each sample when applied to nonlinear prediction problems, and thus
consisted of the microstructure features extracted for a sub- it suits this application. Other methods such as the scaled
image and the corresponding mechanical properties. The conjugate gradient algorithm and resilient backpropaga-
major advantage of this sub-imaging process is to generate tion algorithm were tested, but none was as effective as the
a sufficient number of datasets to train the neural networks Levenberg–Marquardt algorithm. In tests conducted with 50
without requiring a very large number of image property nodes, the LM algorithm far outpaced each of the alterna-
datasets, which could be very expensive and yet generate tives, as demonstrated in Table 4. The next most effective
necessary persistent excitation to warrant good training. was the resilient backpropagation algorithm, though none
The predictive models in this work were built using sin- was able to consistently produce a correlation coefficient
gle-hidden layer feedforward neural networks, which could above 0.9 other than the LM algorithm.
learn to approximate any continuous nonlinear mapping After experimenting with the performance of UTS pre-
from input to output given a sufficient number of hidden diction as the number of hidden nodes was varied, a feed-
nodes. Since each mechanical property of interest would be forward neural network with 55 nodes was used. This net-
predicted, respectively, by an individual model, each neural work achieved a reasonable root-mean-square error (RMSE)
network would have 13 input variables (grain morphologi- without overfitting the data. Tests on the network were con-
cal features) and a single output variable (the mechanical ducted between 5 and 140 nodes at 5 node increments, and
property). Eleven of these inputs were obtained from grain performance was found to be stable between about 45 and
size histograms. As previously described, the histogram bin 90 nodes when using the Levenberg–Marquardt algorithm
values described the proportion of grains in a sample that as shown in Fig. 11. Ultimately, a value of 55 nodes was

Table 3  Summary of neural Variable Avg. level Std Range


network input and output
variables Input Probability in grain size bin#1 [0.31, 0.56] µm 4.62E−03 1.36E−02 [0, 1.30E−01]
Probability in grain size bin#2 [0.56, 1] µm 4.14E−02 8.43E−02 [0, 6.02E−01]
Probability in grain size bin#3 [1, 1.78] µm 9.64E−02 1.39E−01 [0, 6.48E−01]
Probability in grain size bin#4 [1.78, 3.16] µm 1.21E−01 1.33E−01 [0, 5.41E−01]
Probability in grain size bin#5 [3.16, 5.62] µm 1.60E−01 9.77E−02 [0, 6.06E−01]
Probability in grain size bin#6 [5.62, 10] µm 2.26E−01 1.47E−01 [0, 6.16E−01]
Probability in grain size bin#7 [10, 17.78] µm 2.05E−01 1.41E−01 [0, 6.21E−01]
Probability in grain size bin#8 [17.78, 31.62] µm 1.06E−01 1.01E−01 [0, 6.15E−01]
Probability in grain size bin#9 [31.62, 56.23] µm 3.59E−02 5.67E−02 [0, 4.26E−01]
Probability in grain size bin#10 [56.23, 100] µm 3.10E−03 8.92E−03 [0, 1.06E−01]
Probability in grain size bin#11 > 100 µm 1.95E−04 1.52E−03 [0, 3.33E−02]
Mean aspect ratio 1.96E+00 2.10E−01 [1.45E+00, 3.19E+00]
Mean perimeter to diameter ratio 3.79E+00 4.07E−01 [2.60E+00, 6.07E+00]
Output UTS 9.64E+02 1.26E+02 [7.44E+02, 1.41E+03]
Yield strength 7.53E+02 1.35E+02 [5.00E+02, 1.26E+03]
Fracture strain 6.34E−02 3.06E−02 [3.70E−02, 1.21E−01]

13
250 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

Table 4  Algorithm performance Algorithm 5 Nodes 10 Nodes 25 Nodes 50 Nodes 75 Nodes 100 Nodes
at selected node counts (full-
image validation R2) Levenberg–Marquardt 0.882 0.928 0.928 0.946 0.967 0.931
BFGS quasi-Newton 0.774 0.715 0.799 0.789 0.813 0.830
Resilient backpropagation 0.786 0.812 0.825 0.871 0.879 0.898
Scaled conjugate gradient 0.665 0.744 0.658 0.860 0.880 0.844
One-step secant 0.637 0.685 0.850 0.758 0.870 0.654

Fig. 11  Network performance using 5–140 nodes

selected as it was near the lower end of the reliably consist- divided by physical samples, which means if a sample built
ent, stable region. This meant that using a value of 55 nodes in “Physical sample preparation” section was selected for
produced consistent results without the potential danger of validation, then all the sub-image data points originating
overfitting the model and harming performance. The metric from this physical sample would be put into the validation
used to evaluate the number of nodes to be used was the R2 set. Thereby, the trained model will be validated on data
correlation between the predicted values and the UTS value points from additively manufactured samples that have
of the full images. Ultimately, the network’s performance been never seen by the training routine. The data points
improved between this stage of development and when final from about 60% of physical samples were used for training,
results were collected, but Fig. 11 contains some early per- and the data from the remaining 40% samples were used for
formance data that helped to motivate the selection of 55 validation.
nodes. The network size selected for UTS prediction was The prediction results of UTS by the trained neural net-
also applied to the networks predicting other mechanical work are shown in Fig. 11 and Table 4. While the network
properties, as they are all based on the same dataset and was trained by taking each sub-image as an individual data
microstructure features. point, its predictions were assessed in an aggregated way
because the mechanical properties should be more depend-
ent on the overall microstructure instead of the local ones in
Results and Discussion sub-images. Two aggregation methods were employed: One
is to collect the predictions for all the cropped sub-images
With the selected neural network structure, extracted fea- belonging to the same physical sample and then calculate the
tures, and measured (or extracted from the literature) mean and standard deviation of sub-image predictions, and
mechanical properties, three predictive models for the the other is to directly use the predictions based on the fea-
ultimate tensile strength (UTS), yield strength (YS), and tures extracted from the original full images without crop-
fracture strain were trained, respectively. The dataset aug- ping. As can be seen from Table 5, both the mean predic-
mented with sub-image cropping was divided into training tions of sub-image and the full-image predictions achieved
and validation sets. To prevent data leakage, the dataset was excellent adjusted R2, root-mean-square error (RMSE), and

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 251

Table 5  Predictive model Sub-image mean Full image


performance for ultimate tensile
2
strength Adjusted R RMSE (MPa) MAPE (%) Adjusted R2 RMSE (MPa) MAPE

Training 0.988 18.8 1.24 0.982 23.1 1.81


Validation N/A N/A 2.60 0.949 30.9 2.67

mean absolute percentage error (MAPE). For the valida- In the stress–strain curve, locating the yield point was
tion samples, only the full-image predictions were imple- more difficult than the UTS, as the latter was simply the
mented. The prediction accuracy on the validation samples highest point of the true stress curve. Nevertheless, the
not used for training, though degraded slightly compared to MAPE on the validation set was approximately 3.4% of the
the training samples, was also quite good, which proved that yield strength magnitude, which still provides sufficient
the UTS model built in this work does not only possess high accuracy to discern microstructures with different yield
prediction accuracy but also exhibits good generalizability strengths and assess the quality of AM fabricated parts.
to new data. In addition, it can be seen from Fig. 12 that the Lastly, the prediction of fracture strain is shown in
confidence intervals formed based on the standard deviation Fig. 14 and Table 7. The fracture strain indicates when the
of sub-image predictions always enclosed the true measured material breaks under tensile stress and hence measures
UTS values (the solid black line), which provided a quantifi- the ductility of the material. The fracture strain prediction
cation of uncertainties associated with the UTS model. It is exhibited good accuracy on both training and validation
worth noting that for some samples, the confidence intervals sets. Its R2 was better than that of yield strength, while its
appeared to be wide. This is mainly because the microstruc- MAPE was moderately larger. More importantly, similar
tures of these samples were inhomogeneous and thus their to the UTS prediction, the prediction accuracies of both
cropped sub-images exhibited variant grain morphologies, yield strength and fracture strain were fairly consistent
which made the sub-image predictions scattered. Neverthe- among training and validation sets. Considering that the
less, the mean predictions of sub-images were less sensitive validation sets consisted of physical samples whose sub-
to local variants and hence always stayed close to the true images were all excluded from training, this implies that
UTS values. This is why the sub-image predictions should the predictive models of UTS, yield strength, and fracture
be aggregated to report the overall mechanical properties. strain can be reliably generalized to new AM samples of
Upon the success of UTS prediction, the same neu- the 17-4H material. The exemplary generalizability was
ral network structure and training method were applied achieved because the microstructure features, such as the
to the prediction of yield strength. As shown in Fig. 13 grain size histogram inspired by the Hall–Petch relation-
and Table 6, the performance of yield strength prediction ship and the grain aspect ratio to capture the elongated
was still fairly good, though it degraded slightly compared grains in the martensite phase, embraced the first-principle
with the UTS prediction. The lower prediction accuracy knowledge of material mechanics. Therefore, the neural
of yield strength is probably because the yield strength networks in this work can be seen as data-driven fittings
measurement to label training samples was less accurate. of the generally applicable governing physics of materials,

Fig. 12  Neural network prediction for ultimate tensile strength

13
252 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

Fig. 13  Neural network prediction for yield strength

Table 6  Predictive model Sub-image mean Full image


performance for yield strength
2
Adjusted R RMSE (MPa) MAPE (%) Adjusted R2 RMSE (MPa) MAPE (%)

Training 0.973 28.6 2.92 0.969 30.6 3.06


Validation N/A N/A 3.23 0.924 33.8 3.37

Fig. 14  Neural network prediction for fracture strain

Table 7  Predictive model Sub-image mean Full image


performance for fracture strain
2
Adjusted R RMSE MAPE (%) Adjusted R2 RMSE MAPE (%)

Training 0.980 0.0044 4.70 0.967 0.0057 6.93


Validation N/A N/A 8.17 0.949 0.0068 7.44

which are less prone to overfitting than the models con- supervised machine learning techniques available for pre-
structed from a pure data perspective. dictive regression problems. Several alternative modeling
While the above results are all based on feedforward neu- techniques were tested and are compared in Table 8. All the
ral networks with a single hidden layer, there are numerous tests were for the prediction of UTS, and the RMSE reported

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 253

Table 8  Performance of Model structure RMSE on validation


alternative models dataset (UTS data)
(MPa)

Baseline: neural network with one hidden layer (55 nodes) 41.70
Linear regression 87.25
Support vector machine with linear kernel 91.39
Support vector machine with Gaussian kernel 63.30
Support vector machine with cubic kernel 70.81
Random forest (size = 50) 44.63
Random forest (size = 100) 41.96
Neural network with two hidden layers (30/30 nodes) 41.61
Neural network with three hidden layers (20/20/20 nodes) 41.19

in Table 8 was evaluated using the sub-image data samples data point), and it is a signed quantity whose magnitude of
in the validation datasets without applying the aggrega- contribution to the prediction of that query point is indicated
tion strategy, which is why the validation RMSE reported by its absolute value. In Table 9, the absolute Shapley values
for the baseline model was a bit higher than that with averaged over all the full-image samples are presented to
full-image aggregation in Table 4. All the trainings were compare the overall importance of the microstructure fea-
repeated 10 times, and the average of RMSE was reported tures for each of the three mechanical properties. A larger
for each model. As can be seen, the linear regression model Shapley value in Table 9 indicates the feature that has a
performed significantly worse than other machine learn- higher impact on the prediction of corresponding mechani-
ing models, because it could not capture the nonlinearity cal property. The three most important features for each
between input variables and the final material properties. mechanical property are also highlighted in Table 9. As
Support vector machine (SVM) modeling, no matter what expected, the probabilities in grain size bins with lower grain
type of kernel was used, was also unsuccessful compared to diameter and the grain aspect ratio feature had higher contri-
neural networks, which indicates that SVM modeling was butions. As shown in Fig. 9, most of the grains fall into the
not a suitable strategy for this application. For neural net- 1 to 17.78 µm range and the Hall–Petch curve is steeper in
works with two or three hidden layers, given that their num- the low grain diameter range, and, therefore, the mechani-
bers of hidden nodes were both approximately the same as cal properties should be more sensitive to the probabilities
the baseline model, their performance did not exhibit notice- in the bins from 1 to 17.78 µm. In addition, the grain aspect
able improvement by making the network deeper. Therefore, ratio feature implicitly encodes phase information into neu-
the neural network with single hidden layer (i.e., the baseline ral network models (i.e., a higher fraction of martensite
model) was preferable owing to its simpler structure. Finally, phase yields a higher grain aspect ratio, higher strength, and
random forest models were also shown to be powerful in lower fracture strain). Therefore, it can be concluded that the
accurately predicting the UTS, as the random forest with observation of feature relative importance matches with the
100 trees achieved the same level of RMSE as the baseline expectation based on the knowledge of material mechanics,
model. Both the random forest and neural network strate- which is another advantage of using interpretable physics-
gies were successful when applied to this dataset. Given inspired features in data-driven predictive models.
the nature of the present data, it makes sense that a random
forest could be an applicable model. That said, in the future,
as more comprehensive image data could be included in the
model with more complex relations between a larger number Conclusions
of variables, a neural network would likely prove to be more
scalable. In this paper, data-driven models based on microstructure
To further evaluate the importance of adopted microstruc- attributes were established to predict the mechanical prop-
ture features, the Shapley values were calculated for each erties of 17-4H stainless steel material built by additive
feature and the neural network model of each mechanical manufacturing. Predictions were based upon data from three
property. The Shapley value is a metric in game theory to sources: in-house tensile testing and imaging, available lit-
evaluate the relative contribution of each input predictor fea- erature, and additive manufacturing company, Digital Metal
ture to the overall prediction [24]. The Shapley value is typi- and Innovative 3D Manufacturing. By utilizing grain size
cally evaluated for a query point (i.e., each microstructure and shape characteristics derived from images, ANNs were

13
254 Integrating Materials and Manufacturing Innovation (2022) 11:241–255

Table 9  Evaluation of the Input feature Shapley value Shapley value Shapley value
importance of input features to for UTS (MPa) for yield strength for fracture
neural networks (MPa) strain

Probability in grain size bin#1 [0.31, 0.56] µm 5.160127 25.00539 1.61E−03


Probability in grain size bin#2 [0.56, 1] µm 23.21459 0.311903 3.75E−03
Probability in grain size bin#3 [1, 1.78] µm 35.36107 13.56977 1.07E−02
Probability in grain size bin#4 [1.78, 3.16] µm 62.60749 6.220818 1.03E−02
Probability in grain size bin#5 [3.16, 5.62] µm 19.8183 5.504887 6.56E−03
Probability in grain size bin#6 [5.62, 10] µm 27.9357 3.328806 9.01E−03
Probability in grain size bin#7 [10, 17.78] µm 17.73263 23.57832 2.06E−02
Probability in grain size bin#8 [17.78, 31.62] µm 19.65028 17.67157 5.24E−03
Probability in grain size bin#9 [31.62, 56.23] µm 11.36135 15.84337 4.47E−03
Probability in grain size bin#10 [56.23, 100] µm 1.875369 4.515326 3.59E−04
Probability in grain size bin#11 > 100 µm 1.427833 0.490937 3.26E−04
Mean aspect ratio 21.67412 33.6412 2.32E−03
Mean perimeter-to-diameter ratio 7.51172 9.634381 5.06E−03

Absolute Shapley values averaged for all the full-image samples, larger values indicating higher contribu-
tion to ANN predictions. Bold: the top three features for each mechanical property

utilized to derive relationships between said characteristics References


and tensile properties observed in testing.
Results were encouraging across three different tensile 1. Tenner J, Linkens DA, Morris PF, Bailey TJ (2001) Prediction
properties. For ultimate tensile strength, the model was able of mechanical properties in steel heat treatment process using
neural networks. Ironmak Steelmak 28(1):15–22. https://​doi.​org/​
to predict within an average error of 30 MPa of the true 10.​1179/​irs.​2001.​28.1.​15
value (3.1% of mean true UTS value). For yield strength, 2. Sterjovski Z, Nolan D, Carpenter KR, Dunne DP, Norrish J (2005)
this prediction was on average within 33 MPa of the true Artificial neural networks for modelling the mechanical proper-
value (4.5% of the mean true YS value). Finally, for fracture ties of steels in various applications. J Mater Process Technol
170(3):536–544. https://d​ oi.o​ rg/1​ 0.1​ 016/j.j​ matpr​ otec.2​ 005.0​ 5.0​ 40
strength, predicted values matched observed values within 3. DebRoy T, Mukherjee T, Wei HL, Elmer JW, Milewski JO (2020)
an average of 0.7% elongation in comparison with true val- Metallurgy, mechanistic models and machine learning in metal
ues (11% magnitude of true FS). Also, all predictions fell printing. Nat Rev Mater 6(1):48–68. https://​doi.​org/​10.​1038/​
within quantified confidence intervals. These results indi- s41578-​020-​00236-1
4. Popova E, Rodgers TM, Gong X, Cecen A, Madison JD, Kalidindi
cated that all the established predictive models of mechani- SR (2017) Process-structure linkages using a data science
cal properties exhibited practical accuracy and reliability. approach: application to simulated additive manufacturing data.
The key strength of this approach is the model’s general- Integr Mater Manuf Innov 6(1):54–68. https://​doi.​org/​10.​1007/​
izability. By forming linkages between microstructures and s40192-​017-​0088-1
5. Herriot C, Spear AD (2020) Predicting microstructure-dependent
properties, the process utilized in part manufacturing can mechanical properties in additively manufactured metals with
be generalized. As such, this method can incorporate data machine- and deep-learning methods. Comput Mater Sci. https://​
from samples built by many methods, both AM and tradi- doi.​org/​10.​1016/j.​comma​tsci.​2020.​109599
tional, and the inclusion of a wider variety of manufacturing 6. Wan HY, Chen GF, Li CP, Qi XB, Zhang GP (2019) Data-driven
evaluation of fatigue performance of additive manufactured parts
processes could even enhance the accuracy and reliability using miniature specimens. J Mater Sci Technol 35(6):1137–1146.
of its predictions. This procedure could also be adapted to a https://​doi.​org/​10.​1016/j.​jmst.​2018.​12.​011
wider variety of metals for similar results or exposed to data 7. Yang Z, Yu C-H, Buehler MJ (2021) Deep learning model to
from more sources to improve upon predictive capabilities. predict complex stress and strain fields in hierarchical composites.
Sci Adv. https://​doi.​org/​10.​1126/​sciadv.​abd74​16
8. Bayraktar Ö, Uzun G, Çakiroğlu R, Guldas A (2017) Experi-
Acknowledgements The authors gratefully acknowledge the tensile mental study on the 3D-printed plastic parts and predicting the
specimens and microstructural images provided by the Digital Metals mechanical properties using artificial neural networks. Polym Adv
and Innovative 3D Manufacturing during this study. During this study, Technol 28(8):1044–1051. https://​doi.​org/​10.​1002/​pat.​3960
all the experimental work was supported by the Donald A. & Nancy G. 9. Xie X, Bennett J, Saha S, Lu Y, Cao J, Liu WK, Gan Z (2021)
Roach Professorship at Purdue University. Mechanistic data-driven prediction of as-built mechanical proper-
ties in metal additive manufacturing. npj Comput Mater. https://​
Declarations doi.​org/​10.​1038/​s41524-​021-​00555-z
10. Lu L, Dao M, Kumar P, Ramamurty U, Karniadakis GE, Suresh S
Conflict of interest The authors have no conflict of interest. (2020) Extraction of mechanical properties of materials through

13
Integrating Materials and Manufacturing Innovation (2022) 11:241–255 255

deep learning from instrumented indentation. Proc Natl Acad Sci melting process induced oxide dispersion strengthened 17-4 PH
117(13):7052–7062. https://​doi.​org/​10.​1073/​pnas.​19222​10117 stainless steel. J Alloy Compd 803(1):30–41. https://​doi.​org/​10.​
11. Yan W, Lian Y, Yu C, Kafka OL, Liu Z, Liu WK, Wagner GJ 1016/j.​jallc​om.​2019.​06.​289
(2018) An integrated process–structure–property modeling frame- 18. Mahmoudi M, Elwany A, Yadollahi A, Thompson SM, Bian L,
work for additive manufacturing. Comput Methods Appl Mech Shamsaei N (2017) Mechanical properties and microstructural
Eng 339(1):184–204. https://​doi.​org/​10.​1016/j.​cma.​2018.​05.​004 characterization of selective laser melted 17-4 PH stainless
12. Gao W, Zhang Y, Ramanujan D, Ramani K, Chen Y, Williams CB, steel. Rapid Prototyp J 23(2):280–294. https://​doi.​org/​10.​1108/​
Wang CCL, Shin YC, Zhang S, Zavattieri PD (2015) The status, rpj-​12-​2015-​0192
challenges, and future of additive manufacturing in engineering. 19. Alnajjar M, Christien F, Bosch C, Wolski K (2020) A comparative
Comput-Aided Des 69(1):65–89. https://​doi.​org/​10.​1016/j.​cad.​ study of microstructure and hydrogen embrittlement of selective
2015.​04.​001 laser melted and wrought 17–4 PH stainless steel. Mater Sci Eng
13. Yadollahi A, Shamsaei N, Thompson SM, Elwany A, Bian L A. https://​doi.​org/​10.​1016/j.​msea.​2020.​139363
(2015) mechanical and microstructural properties of selective 20. Yadollahi A, Shamsaei N, Thompson SM, Elwany A, Bian L
laser melted 17-4 PH stainless steel. In: Proceedings of the ASME (2017) Effects of building orientation and heat treatment on
2015 international mechanical engineering congress and exposi- fatigue behavior of selective laser melted 17–4 PH stainless steel.
tion, 2A(1). https://​doi.​org/​10.​1115/​imece​2015-​52362 Int J Fatigue 94(1):218–235. https://​doi.​org/​10.​1016/j.​ijfat​igue.​
14. Rafi HK, Pal D, Patil N, Starr TL, Stucker BE (2014) Microstruc- 2016.​03.​014
ture and mechanical behavior of 17-4 precipitation hardenable 21. Wang X, Liu Y, Shi T, Wang Y (2020) Strain rate dependence of
steel processed by selective laser melting. J Mater Eng Perform mechanical property in a selective laser melted 17–4 PH stainless
23(12):4421–4428. https://​doi.​org/​10.​1007/​s11665-​014-​1226-y steel with different states. Mater Sci Eng A. https://​doi.​org/​10.​
15. Yadollahi A, Mahmoudi M, Elwany A, Doude H, Bian L, New- 1016/j.​msea.​2020.​139776
man JC (2020) Effects of crack orientation and heat treatment on 22. Callister WD (2000) Fundamentals of materials science and engi-
fatigue-crack-growth behavior of AM 17-4 PH stainless steel. Eng neering, vol 471660817. Wiley, London
Fract Mech. https://​doi.​org/​10.​1016/j.​engfr​acmech.​2020.​106874 23. Raeisinia B, Poole WJ (2011) Modelling the elastic–plastic transi-
16. Pasebani S, Ghayoor M, Badwe S, Irrinki H, Atre SV (2018) tion of polycrystalline metals with a distribution of grain sizes.
Effects of atomizing media and post processing on mechanical Modell Simul Mater Sci Eng 20(1):015015. https://​doi.​org/​10.​
properties of 17-4 PH stainless steel manufactured via selective 1088/​0965-​0393/​20/1/​015015
laser melting. Addit Manuf 22(1):127–137. https://​doi.​org/​10.​ 24. Kumar IE, Venkatasubramanian S, Scheidegger C, Friedler S
1016/j.​addma.​2018.​05.​011 (2020) Problems with Shapley-value-based explanations as feature
17. Hsu TH, Chang YJ, Huang CY, Yen HW, Chen CP, Jen KK, importance measures. In: International conference on machine
Yeh AC (2019) Microstructure and property of a selective laser learning. PMLR, pp 5491–5500

13

View publication stats

You might also like