Applied Sciences: Machine Learning in Manufacturing Towards Industry 4.0: From For Now' To Four-Know'

Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

applied

sciences
Article
Machine Learning in Manufacturing towards Industry 4.0:
From ‘For Now’ to ‘Four-Know’
Tingting Chen 1, * , Vignesh Sampath 2 , Marvin Carl May 3 , Shuo Shan 1 , Oliver Jonas Jorg 4 ,
Juan José Aguilar Martín 5 , Florian Stamer 3 , Gualtiero Fantoni 4 , Guido Tosello 1 and Matteo Calaon 1

1 Department of Civil and Mechanical Engineering, Technical University of Denmark,


2800 Kongens Lyngby, Denmark
2 Autonomous and Intelligent Systems Unit, Tekniker, Member of Basque Research and Technology Alliance,
20600 Eibar, Spain
3 wbk Institute of Production Science, Karlsruhe Institute of Technology (KIT), Kaiserstr. 12,
76131 Karlsruhe, Germany
4 Department of Civil and Industrial Engineering, University of Pisa, 56122 Pisa, Italy
5 Department of Design and Manufacturing Engineering, School of Engineering and Architecture,
University of Zarazoga, 50009 Zaragoza, Spain
* Correspondence: [email protected]; Tel.: +45-5028-1068

Abstract: While attracting increasing research attention in science and technology, Machine Learning
(ML) is playing a critical role in the digitalization of manufacturing operations towards Industry 4.0.
Recently, ML has been applied in several fields of production engineering to solve a variety of tasks
with different levels of complexity and performance. However, in spite of the enormous number
of ML use cases, there is no guidance or standard for developing ML solutions from ideation to
deployment. This paper aims to address this problem by proposing an ML application roadmap for
the manufacturing industry based on the state-of-the-art published research on the topic. First, this
paper presents two dimensions for formulating ML tasks, namely, ’Four-Know’ (Know-what, Know-
why, Know-when, Know-how) and ’Four-Level’ (Product, Process, Machine, System). These are used
Citation: Chen, T.; Sampath, V.; May, to analyze ML development trends in manufacturing. Then, the paper provides an implementation
M.C.; Shan, S.; Jorg, O.J.; Aguilar pipeline starting from the very early stages of ML solution development and summarizes the available
Martín, J.J.; Stamer, F.; Fantoni, G.; ML methods, including supervised learning methods, semi-supervised methods, unsupervised
Tosello, G.; Calaon, M. Machine methods, and reinforcement methods, along with their typical applications. Finally, the paper
Learning in Manufacturing towards discusses the current challenges during ML applications and provides an outline of possible directions
Industry 4.0: From ‘For Now’ to for future developments.
‘Four-Know’. Appl. Sci. 2023, 13, 1903.
https://doi.org/10.3390/app13031903
Keywords: machine learning; Industry 4.0; manufacturing; artificial intelligence; smart manufacturing;
Academic Editors: Alexandre digitization
Carvalho and Richard (Chunhui)
Yang

Received: 23 November 2022


1. Introduction
Revised: 18 January 2023
Accepted: 27 January 2023 Within the fourth industrial revolution, coined as ‘Industry 4.0’, the way products
Published: 1 February 2023 are manufactured is changing dramatically [1]. Moreover, the way humans and machines
interact with one another in manufacturing has seen enormous changes [2], developing
towards an ‘Industry 5.0’ notion [3]. The digitalization of businesses and production compa-
nies, the inter-connection of their machines through embedded system and the Internet of
Copyright: © 2023 by the authors. Things (IoT) [4], the rise of cobots [5,6], and the use of individual workstations and matrix
Licensee MDPI, Basel, Switzerland. production [7] are disrupting conventional manufacturing paradigms [1,8]. The demand for
This article is an open access article
individualized and customized products is continuously increasing. Consequently, order
distributed under the terms and
numbers are surging while batch sizes diminish, to the extremes of fully decentralized
conditions of the Creative Commons
’batch size one’ production. The demand for a high level of variability in production and
Attribution (CC BY) license (https://
manufacturing through Mass Customization is inevitable. Mass Customization in turn
creativecommons.org/licenses/by/
requires manufacturing systems which are increasingly more flexible and adaptable [7–9].
4.0/).

Appl. Sci. 2023, 13, 1903. https://doi.org/10.3390/app13031903 https://www.mdpi.com/journal/applsci


Appl. Sci. 2023, 13, 1903 2 of 32

Machine Learning (ML) is one of the cornerstones for making manufacturing (more)
intelligent, and thereby providing it with the needed capabilities towards greater flexibility
and adaptability [10]. These advances in ML are shifting the traditional manufacturing era
into the smart manufacturing era of Industry 4.0 [11]. Therefore, ML plays an increasingly
important role in manufacturing domain together with digital solutions and advanced
technologies, including the Industrial Internet of Things (IIoT), additive manufacturing,
digital twins, advanced robotics, cloud computing, and augmented/virtual reality [11].
ML refers to a field of Artificial Intelligence (AI) that covers algorithms learning directly
from their input data [12]. Despite most researchers focusing on finding a single suitable
ML solution for a specific problem, efforts have already been undertaken to reveal the
entire scope of ML in manufacturing. Wang et al. presented frequently-used deep learning
algorithms along with an assessment of their applications towards making manufacturing
“smart” in their 2018 survey [13]. In particular, they discussed four learning models:
Convolutional Neural Networks, Restricted Boltzmann Machines, Auto-Encoders, and
Recurrent Neural Networks. In their recent literature review on “Machine Learning for
Industrial Applications”, Bertolini et al. [12] identified, classified, and analyzed 147 papers
published during a twenty-year time span from Jan. 2000 to Jan. 2020. In addition, they
provided a classification on the basis of application domains in terms of both industrial
areas and processes, as well as their respective subareas. Within these domains, the
authors analyzed the different trends concerning supervised, unsupervised, and reinforced
learning techniques, including the most commonly used algorithms, Neural Networks
(NNs), Support Vector Machine (SVM), and Tree-Based (TB) techniques. The goal of another
literature review from Dogan and Birant [14] was to provide a sound comprehension of
the major approaches and algorithms from the fields of ML and data mining (DM) that
have been used to improve manufacturing in the recent past. Similarly, they investigated
research articles from the period of the past two decades and grouped the identified articles
under four main subjects: scheduling, monitoring, quality, and failure.
While these classifications and trend analyses provide an excellent overview of the
extent of ML applications in manufacturing, they mainly focus on introducing ML algo-
rithms; the implementation of ML solution for different tasks in an industrial environment
from scratch has not yet been fully discussed. In general, a comprehensive formulation of
industrial problems prior to the development of ML solutions seems lacking. Therefore,
the issue we aim to address in this paper is how ML can be implemented to improve manu-
facturing in the transition towards Industry 4.0. From this issue, we derive the following
research questions:
• RQ1: How does ML benefit manufacturing, and what are the typical ML application
cases?
• RQ2: How are ML-based solutions developed for problems in manufacturing engineering?
• RQ3: What are the challenges and opportunities in applying ML in manufacturing contexts?
To answer these research questions, more than a thousand research articles retrieved
from two well-known research databases were systematically identified, screened, and
analyzed. Subsequently, the articles were classified within a two-dimensional framework,
which takes value-based development stages into account on one axis and manufacturing
levels on the other. The development stage concerns visibility, transparency, predictive
capacity, and adaptability, whereas the four manufacturing levels are product, process,
machine, and system.
The rest of this paper is structured as follows. Section 1 introduces the key concepts,
research questions, and motivations. Section 2 proposes the methodology of ’Four-know’
and ’Four-level’ to establish a two-dimensional framework for helping to formulate indus-
trial problems effectively. Based on the proposed framework, a systematic literature review
is carried out and the identified articles are analysed and classified. Section 3 describes a
six-step pipeline for the application of ML in manufacturing. Section 4 explains different
ML methods, presenting where and how they have been applied in manufacturing accord-
ing to the prior identified research articles. Section 5 formulates common challenges and
Appl. Sci. 2023, 13, 1903 3 of 32

potential future directions; finally, the paper concludes in Section 6 with a summary and
discussion of the authors’ findings.

2. Overview of Machine Learning in Manufacturing


Despite numerous ML studies and their promising performance, it remains very
difficult for non-experts working in the manufacturing industry to begin developing ML
solutions for their specific problems. The first challenging part of application is to formulate
the actual problems to be solved [15]. Therefore, this section aims to overcome this problem
by introducing the categories of Four-Know and Four-Level to help formulate ML tasks in
manufacturing and describing the benefits of applying ML in manufacturing from ML use
cases categorized using the Four-Know and Four-Level concepts (RQ1). Lastly, an overview
and developing trends in recent ML studies are provided as formulated by Four-Know and
Four-Level.

2.1. Introduction of Four-Know and Four-Level


According to the Acatech Industrie 4.0 Maturity levels [16], the development towards
Industry 4.0 in manufacturing can be structured into the following six successive stages:
computerization, connectivity, visibility, transparency, predictive capacity, and adaptability.
The first two stages, computerization and connectivity, provide the basis for digitization,
while the rest are analytic capabilities required for achieving Industry 4.0. ML, as powerful
data analytics tools are normally applied in the last four stages. Inspired by the Acatech
Industrie 4.0 Maturity levels, ML studies in manufacturing can be categorized into four
subjects: Know-what, Know-why, Know-when, and Know-how, which to a degree overlap
with visibility, transparency, predictive capacity, and adaptability, respectively. The Four-
Know definitions are presented below:
• Know-what deals with understanding of the current states of machines, processes, or
production systems, which can help in rapid decision-making. It should be noted
that Know-what goes beyond visualization of real-time data. Instead, data should be
processed, analyzed, and distilled into information which enables decision-making.
For instance, typical examples of Know-what in manufacturing are defect detection
in quality control [17,18], fault detection in process/machine monitoring [19,20], and
soft sensor modelling [21,22].
• Know-why, based on the information from Know-what, aims to identify inner patterns
from historical data, thereby discovering the reasons for a thing happening. Know-
why includes the identification of interactions among different variables [23] and the
discovery of cause-effect relationship between an event and other variables [24,25].
On one hand, Know-why can indicate most important factors for understanding
Know-what. On the other hand, Know-why is the prerequisite for Know-when, as the
reliability of predictions is heavily dependent upon the quality of casual inference.
• Know-when, built on Know-why, involves timely predictions of events or prediction
of key variables based on historical data, allowing the decision-maker can take ac-
tions at early stages. For instance, Know-when in manufacturing includes quality
prediction based on relevant variables [26,27], predictive maintenance via detection of
incipient anomalies before break-down [28,29], and predicting Remaining Useful Life
(RUL) [30,31].
• Know-how, on the foundation of Know-when, can recommend decisions that help adapt
to expected disturbance and can aid in self-optimization. Examples in manufacturing
include prediction-based process control [27,32], scheduling of predictive maintenance
tasks [33,34], dynamic scheduling in the flexible production [35,36], and inventory
control [34].
The aim of applying ML in manufacturing is to achieve production optimization across
four different levels: product, process, machine, and system. Therefore, the use cases for
applying ML can be further categorized by these different levels, as shown in Figure 1 and
Table 1, which answer RQ1 in terms of ML typical use cases.
Appl. Sci. 2023, 13, 1903 4 of 32

Table 1. Typical ML use cases categorized by Four-Level and Four-Know.

Level Know-What Know-Why Know-When Know-How


Defect detection [37], Correlation between process Quality improve-
Product Quality prediction [26]
Product design [38] and quality [23] ment [39]
Root cause analysis of pro- Process fault predic- Self-optimizing process
Process Process monitoring [40] cess failure [41], Process tion [43], Process character- planning [45], Adaptive
modelling [42] istics prediction [44] process control [46]
Machine tool monitor- Fault diagnosis [48], Down- RUL prediction [50], Tool Adaptive compensation
Machine
ing [47] time prediction [49] wear prediction [51] of errors [52,53],
Root cause analysis of
Production performance Predictive schedul-
production disturbances
System Anomaly detection [54] prediction [56], Human ing [58], Adaptive
or casual-relationship
behavior control [57] production control [59]
discovery [55]

Figure 1. Four-Level and Four-Know categorization of ML applications. The Four-Know categories,


from Know-what to Know-how, are respectively demonstrated by the four concentric circles, from the
inner circle to the outer circle, with each circle divided into four quarters according to the Four Levels.

2.2. Literature Review Methodology


In order to address the research questions laid out in Section 1, a systematic litera-
ture review following the PRISMA methodology [60] was carried out. Two well-known
research databases, Scopus (Elsevier) and Web of Science (WoS), were chosen for retrieving
documents. The overall literature review process is shown in Figure 2.
Appl. Sci. 2023, 13, 1903 5 of 32

Figure 2. The overall literature review process following PRISMA. All identified documents were
screened and assessed for eligibility, then subjected to Four-Level and Four-Know classification.

Table 2 shows the limitations used when performing the document search. It should
be noted that the query strings were used for Title, Abstract, and Keywords as well as
Keyword Plus (only in WoS).

Table 2. Limitations for document searching.

Item Description
( “manufacturing” OR “industry 4.0” OR “industrie 4.0” )
AND ( “machine learning” OR “deep learning” OR “super-
Query string
vised learning” OR “semi-supervised learning” OR “unsuper-
vised learning” OR “reinforcement learning” )
Year Published from 2018 to 2022
Language English
Subject/Research area Engineering
Document type Article

Following the document search, 2547 documents were found from Scopus and 1784
from WoS. The identified publications from the two databases were merged and duplicates
were removed, resulting in 2861 publications. The documents were then evaluated and
selected by reading the Title and Abstract field, and articles that did not meet the following
selection criteria were excluded:
• The study dealt with the context of manufacturing;
• The study dealt with ML applications in specific fields.
Therefore, conceptual models, frameworks, and studies that only focused on algorithm
development were considered to be out of scope.
Finally, the remaining 1348 documents were analyzed and classified based on the
Four-Level and Four-Know categories. Figure 3 shows the trend of ML applications in
manufacturing over the past five years from the Four-Level perspective. Figure 4 reveals
the detailed distribution of ML applications in Four-Know terms. It should be noted that
because the literature review was conducted in August 2022, the actual numbers for the
Appl. Sci. 2023, 13, 1903 6 of 32

full year 2022 should be higher. As can be seen, there has been a gradual increase in the
number of ML publications in manufacturing in all levels over the past five years. Typically,
what stands out in this figure is the dominance of the product level. From Figure 4, it can
be seen that recent ML applications in product level are mainly focused on Know-what
and Know-when. A similar pattern can be found at the machine level. Interestingly, a
considerable growth in Know-how is observed at the process and system levels compared
to the others. The reason for this may be correlated with higher demand for adaptability
with respect to changes on the process and system levels.
The identified documents were analyzed and classified according to their applied ML
methods, providing examples for non-experts when dealing with similar tasks.

Figure 3. Trends in ML publications in manufacturing in the past five years by Four-Level grouping.

Figure 4. Four-Know development trends for each level over the past five years.

3. Pipeline of Applying Machine Learning in Manufacturing


ML is a technique capable of extracting knowledge from data automatically [12]. In-
creasing research on ML has shown that it is an appealing solution when tackling complex
Appl. Sci. 2023, 13, 1903 7 of 32

challenges. In recent years, more and more manufacturing industries have begun to lever-
age the benefits of ML by developing ML solutions in several industrial fields. However,
despite plenty of off-the-shelf ML models, there are challenges when applying ML to
real-world problems [61]. In particular, it is harder for small and medium-sized enterprises
to develop in-house ML solutions, as commercial ML solutions are normally confidential
and inaccessible. Therefore, this section aims to provide a pipeline for applying ML for
those who are starting from scratch (RQ2). Applying machine learning in manufacturing
normally involves the following six steps: (i) data collection, (ii) data cleaning, (iii) data
transformation, (iv) model training, (v) model analysis, and (vi) model push, as shown in
Figure 5.

Figure 5. Pipeline of applying machine learning in manufacturing.

3.1. Data Collection


The lifeblood of any machine learning model is data. In order for an ML model to
learn, clean data samples must be continuously fed into system throughout the training
process. When the collected data are highly imbalanced or otherwise inadequate, the
desired task may not be achievable. Data can be collected from different sources, including
machines, processes, or production with the aid of sensors or external databases. In terms
of data types, the data used in machine learning can be generally categorized as follows:
• Image data, matrices of pixels with two or more dimensions, such as gray-scale images
or colored images. Image data can acquired by with vision systems, through data
transformations such as simple concatenation of several one-dimensional vectors
with same length, or by the transformation of images from the spatial domain to the
frequency domain.
• Tabular data organized in a table, where normally one axis represents attributes and
another axis represents observations. Tabular data are typically observed in production
data, where the attributes of events of interest are collected. Though tabular data
share a similar data structure with image data, the latter are more focused on one-
dimensional interaction among attributes, while image data typically stress spatial
interactions in both dimensions.
• Time series data, sequences of one or more attributes over time, with the former corre-
sponding to univariate time series and the latter multivariate time series. In manufac-
turing, time series data are normally acquired with sensors whenever there is a need
for monitoring time flow changes of data.
• Text data, including written documents with words, sentences or paragraphs. Ex-
amples of text data in manufacturing include maintenance reports on machines and
descriptions of unexpected disturbances or events in production.

3.2. Data Cleaning


Real-world industrial data are highly susceptible to noisy, missing, and inconsistent
data due to several factors. Low-quality noisy data can lead to less accurate ML models.
Data cleaning [62] is a crucial step when organizing data into a consistent data structure
across packages, and can improve the quality of the data, leading to more accurate ML
models. It is usually performed as an iterative approach. Methods include filling in missing
values, smoothing noisy data, removing outliers, resolving data inconsistencies, etc.
Appl. Sci. 2023, 13, 1903 8 of 32

3.3. Data Transformation


Data transformation is the process of transforming unstructured raw data into data
better suited for model construction. Data transformation can be broadly classified into
mandatory transformations and optional quality transformations. Mandatory transforma-
tions must be carried out to convert the data into a usable format and then deliver the
transformed data to the destination system. These include transforming non-numerical
data into numerical data, resizing data to a fixed size, etc. It should be noted that data
transformations are not always straightforward. Indeed, in certain situations data types
can be interconvertible by leveraging specific processing techniques, as shown in Figure 6.
For instance, univariate time series can be converted into image data using the Gramian
Angular Field (GAF) or Markov Transition Field (MTF) [63] methods. Unstructured text
data can be converted into tabular data via word embedding [64]. Tabular data can be
transformed into image data by projecting data into a 2D space and assigning pixels, as
in Deepinsight [65] or Image Generator for Tabular Data (IGTD) [66]. Image data are
preferable for data analysis, as they allow the power of Convolutional Neural Networks
(CNNs) [67] to be exploited.
In real-world applications, data are normally high-dimensional and redundant. When
performing data modelling directly in the original high-dimensional space, the computa-
tional efficiency can be very low. Hence, it is necessary to reduce the dimensionality in order
to obtain better representation for data modelling. This is achieved by feature selection,
which selects the most informative feature subset from raw data, or feature extraction,
which generates new lower-dimensional features. After feature engineering, features are
either manually designed, so-called “handcrafted features” [68], or automatically learned
from data, so-called “automatic features”. Handcrafted features are heavily dependent on
domain knowledge, and normally have physical meaning. However, these features are
highly subjective [69] and inevitably lack implicit key features [70,71].
By contrast, automatic features driven by data require no prior knowledge. Therefore,
they have been gaining increasing research attention in recent years. Conventionally,
automatic features are obtained by linear transformations such as Principle Component
Analysis (PCA) [72] or Independent Component Analysis (ICA) [73]. However, with the
development of Artificial Neural Networks (ANNs), direct learning of implicit features has
become possible by optimizing the loss function. Thus, neural networks have gradually
developed into an end-to-end solution where knowledge is directly learned from raw data
without human effort. Typically, CNNs [74] and Recurrent Neural networks (RNNs) [75]
are used for image data and time series data, respectively.
A summary of typical features for different data types can be seen in Table 3.

Table 3. Typical features for different data types.

Data Type Handcrafted Features Automatic Features


Image data LBP [76], SIFT [77], HOG [78] ICA, CNNs
Tabular data feature selection PCA, ICA, ANNs
Time domain: mean, min, max, etc.
Time series data ICA, RNNs
Frequency domain: power spectrum [78]
Time-frequency domain: DWT [79], STFT [80]
Text data Bag of Words (BoW) [81] Word2vec [82]

3.4. Model Training


After selecting the features, it is necessary to form the correct data structure for each
individual ML model used in the subsequent steps. Note that different ML algorithms
might require different data models for the same task. Furthermore, results can be improved
through normalization or standardization. Then, the ML models can be applied in the actual
modelling phase. The first step in training a machine learning model typically involves
Appl. Sci. 2023, 13, 1903 9 of 32

selecting a model type that is appropriate for the nature of the data and the problem at
hand. After a model has been chosen, it can be trained by providing it with the training
data and using an optimization algorithm to find the set of parameters that provide the best
performance on those data. Depending on the task, either unsupervised, semi-supervised,
supervised, or reinforcement learning can be applied. These are individually introduced in
the following section.

3.5. Model Analysis


Analysis of model performance is an important step in choosing the right model. This
stage emphasizes how effective the selected model will perform in the future and helps
to make the final decision with regard to model selection. Performance analysis evaluates
models using different metrics, e.g., accuracy, precision, recall, and F1-score (the weighted
average of precision and recall) for classification tasks and the root mean square error
(RMSE) for regression tasks.

3.6. Model Push


Although state-of-the-art ML models improve predictive performance, they contain
millions of parameters, and consequently require a large number of operations per infer-
ence. Such computationally intensive models make deployment in low-power or resource-
constrained devices with strict latency requirements quite difficult. Several methods,
including model pruning [83], model quantization [84], and knowledge distillation [85],
have been suggested in the literature as ways to compress these dense models.
Overall, In the context of manufacturing applications, data collection, data cleaning,
data transformation, model training, model analysis, and model push are key steps in
the implementation of utilizing historical data with ML in order to optimize production
and improve efficiency, quality, and productivity. For instance, data collection involves
gathering data from various sources, such as sensor data, production logs, and quality
control records. Data cleaning involves removing any errors, inconsistencies, or irrelevant
information from the data. Data transformation involves preparing the data for analysis via
formatting in a way that is suitable for the chosen model. Model training involves using the
cleaned and transformed data to train a machine learning model. Model analysis involves
evaluating the performance of the model and identifying any areas for improvement. Model
push involves deploying the model in a production environment and making predictions
or decisions based on the model. All of these steps are critical to ensuring that the results
from ML models are accurate, reliable, and useful for manufacturing production.

Figure 6. Data types used in ML and their convertibility.


Appl. Sci. 2023, 13, 1903 10 of 32

4. Machine Learning Methods and Applications


Model development is the core of ML-based solutions, as the selection of an ML model
plays a critical roles in the outcome. Therefore, this section aims to provide a comprehensive
overview of ML methods and their potential possibilities in manufacturing applications,
including supervised learning methods, semi-supervised learning methods, unsupervised
learning methods, and reinforcement learning methods. In addition, example typical
applications for each category of ML method are listed to support model selection.

4.1. Supervised Learning Methods


Supervised learning methods aim to learn an approximation function f that can map
inputs x to outputs y with the guidance of annotations ( x1 , y1 ), ( x2 , y2 ), . . . , ( x N , y N ). In
supervised learning, the algorithm analyzes a labeled dataset and derives an inferred
function which can be applied to unseen samples. It should be noted that labeled dataset
is a necessity for supervised learning, and as such it requires a large amount of data and
high labeling costs. Supervised learning methods are generally used for dealing with two
problems, namely, regression and classification. The difference between regression and
classification is in the data type of the output variables; regression predicts continuous
numeric values (y ∈ R), while classification predicts categorical values (y ∈ {0, 1}). In
terms of principles, supervised learning methods can be further categorized into four
groups: tree-based methods, probabilistic-based methods, kernel-based methods, and
neural network-based methods.
Tree-based methods: Tree-based methods aim at partitioning the feature space into
several regions until the datapoints in each region share a similar class or value, as depicted
in Figure 7. After space partitioning, a series of if–then rules with a tree-like structure can
be obtained and used to determine the target class or value. Compared with the black-box
models in other supervised methods, Tree-based methods are easily understandable models
that offer better model interpretability. Decision trees [86], in which only a single tree is
established, are the most basic of tree-based methods. It is simple and effective to train a
decision tree, and the results are intuitively understandable, though this approach is very
prone to overfitting. A tree ensemble is an extension of the decision tree concept. Instead of
establishing a single tree, multiple trees are established in parallel or in sequence, referred
to as bagging [87] and boosting [88], respectively. Commonly used tree ensemble methods
include Random Forest [89], Adaptive Boosting (AdaBoost) [88], and Extreme Gradient
Boosting (XGBoost) [90].
Thanks to their better model interpretability, tree-based methods can be used to
identify the most important factors leading up to events. Their possible applications in
manufacturing are mainly in the Know-why and Know-when stages. For instance, exam-
ples of Know-why tasks with tree-based methods at the product and machine level include
identifying the influencing factors that lead to quality defects [91] or machine failure [92],
thereby allowing the manufacturer to diagnose problems effectively. In addition, the identi-
fied important factors when using tree-based methods can help in further predicting target
values such as product quality [93](Know-when, product level) or events of interest before
they happen, such as machine breakdown [31] (Know-when, machine level).
Probabilistic-based methods: For a given input, probabilistic-based methods provide
probabilities for each class as the output. Probabilistic models are able to explain the
uncertainties inherent to data, and can hierarchically build complex models. Widely used
probabilistic-based methods include Bayesian Optimization (BO) [94] and Hidden Markov
Models (HMM) [95].
Appl. Sci. 2023, 13, 1903 11 of 32

Figure 7. The principle of a decision tree. As shown, the feature space is partitioned into several
rectangles in which the input point can find the corresponding class.

The dependencies among different variables can be well captured by Bayesian net-
works [94], enabling a greater likelihood of predicting the target. This can be potentially
beneficial for manufacturing when it comes to Know-what and Know-when tasks, for in-
stance, detection or prediction of events such as quality issues [96] (product level), machine
failure [97] (machine level), or dynamic process modelling [98] (process level).
Markov chains [95], on the other hand, are a type of probabilistic model that describe
a sequence of possible events in which the probability of each event depends only on
the state attained in the previous event. Markov chains can be utilized in manufacturing
to model and analyze the behavior of systems (Know-why, system level) such as pro-
duction lines [99] or supply chains [100]. In addition, the capability of predicting future
states with Markov chains enables applications predicting joint maintenance in production
systems [101] (Know-when, system level) and optimizing production scheduling [102]
(Know-how, system level).
Kernel-based methods: As depicted in Figure 8, kernel-based methods utilize a defined
kernel function to map input data into a high-dimensional implicit feature space [103].
Instead of computing the targeted coordinates, kernel-based methods normally compute
the inner product between a pair of data points in the feature space. However, kernel-based
methods have low efficiency, especially with respect to large-scale input data. Due to the
promising capability of kernel-based methods in classification and regression, they can
be utilized in the Know-what and Know-when stages in manufacturing, such as defect
detection [104] (Know-what, product level), quality prediction [105] (Know-when, product
level), and wear prediction in machinery [106] (Know-when, machine level). There are
different types of kernel-based methods in supervised learning, such as SVM [107] and
Kernel–Fisher discriminant analysis (KFD) [108].

Figure 8. The principle of kernel-based methods. Using a kernel, the linearly inseparable input data
are transformed to another feature space in which they become linearly separable.
Appl. Sci. 2023, 13, 1903 12 of 32

Neural-network-based methods: Inspired by biological neurons and their ability to


communicate with other connected cells, neural network-based methods employ artificial
neurons. A typical neural network, such as ANNs, consists of an input layer, hidden layer,
and output layer, as illustrated in Figure 9. Common ANNs types include CNNs [109],
RNNs [110], and Deep Belief Network (DBN) [111].
Thanks to their powerful feature extraction capability when using matrix-like data,
CNNs are widely used for image processing. In terms of possible applications in man-
ufacturing, CNNs can be used in the Know-what stage to perform image-based quality
control [112] (Know-what, product level) or image-based process monitoring [113] (Know-
what, process level). In addition, by converting time series data from sensors to 2D
images [114], CNNs can be used to detect and diagnosis machine failure as well.
RNNs are typically used to process sequential input data such as time series data or
sequential images. Therefore, in terms of possible applications in manufacturing, RNNs are
well-suited to the Know-when stage for analyzing sensor data or live images from machines,
processes, or production systems. For instance, RNNs can enable the real-time performance
prediction, such as the remaining useful life of machinery [115] (Know-when, machine
level), process behavior prediction [116] (Know-when, process level), or the prediction of
production indicators for real-time production scheduling [117] (Know-when, system level).

Figure 9. The scheme of an ANN, which normally consists of an input layer, hidden layer and
output layer.

The typical supervised learning approaches applied in manufacturing are summarized


in Table A1.

4.2. Unsupervised Learning Methods


Unsupervised learning algorithms aim to identify patterns in data sets containing
data points that are not labeled. Unsupervised learning eliminates the need for labeled
data and manual feature engineering, allowing for more general, flexible, and automated
ML methods. As a result, unsupervised learning methods draw patterns and highlight
areas of interest, revealing critical insight into the production process and opportunities
for improvement. This can allow manufacturers to make better production-focused de-
cisions, driving their business forward. The primary goal of unsupervised learning is to
identify hidden and interesting patterns in unlabeled data. In terms of principles, there
are three types of unsupervised tasks: Dimension Reduction [118,119], Clustering [120],
and Association Rules [121]. Many aspects of unsupervised learning can be beneficial
in manufacturing applications. First, clustering algorithms can be used to identify out-
liers in manufacturing data. Another aspect is to handle high dimensional data, e.g., for
manufacturing cost estimation, quality improvement methodologies, production process
optimization, better understanding of the customer’s data, etc. Usually, a dimensional
reduction support algorithm is required to handle data complexity and high dimensionality.
Finally, it is challenging to perform root cause analysis in large-scale process execution
due to the complexity of services in data centers. Association rule-based learning can be
Appl. Sci. 2023, 13, 1903 13 of 32

employed to conduct root cause analysis and to identify correlations between variables in a
dataset.
Dimensional reduction is the process of converting data from a high-dimensional
space to a low-dimensional space while preserving important characteristics of the
original data.
Principal component analysis (PCA) [118]: The main idea of PCA is to minimize the
number of interrelated variables in a dataset while preserving as much of the dataset’s
inherent variance as possible. A new set of variables, called principal components (PCs),
are generated; these are uncorrelated and sorted such that the first few variables retain the
majority of the variance included in all of the original variables. A pictorial representation
of PCA is shown in Figure 10.

Figure 10. Principal Component Analysis.

The five steps below can be used to condense the entire process of extracting principal
components from a raw dataset.
1. Say we wish to condense d features in our data matrix X to k features. The first step is
to standardize the input data:
z = x − µ/σ
where µ is the mean and σ is the standard deviation.
2. Next, it is necessary to find the covariance matrix of the standardized input data. The
covariance of variables X and Y can be written as follows:
n
1
cov( X, Y ) = ∑
n − 1 i =1
( Xi − ~x )(Yi − ȳ). (1)

3. The third steps is to find all of the eigenvalues and eigenvectors of the covariance matrix:

A~v = λ~v (2)

A~v − λ~v = 0 (3)


~v( A − λI ) = 0. (4)
4. Then, the eigenvector corresponding to the largest eigenvalue is the direction with
the maximum variance, the eigenvector corresponding to second-largest eigenvalue
is the direction with the second maximum variance, etc.
5. To obtain k features, it is necessary to multiply the original data matrix by the matrix
of eigenvectors corresponding to the k largest eigenvalues.
Appl. Sci. 2023, 13, 1903 14 of 32

PCA is particularly useful for processing manufacturing data, which typically have a
large number of variables, making it difficult to identify patterns and trends. A variety of
applications of PCA in manufacturing are listed below:
1. Quality improvement (Know-why, product level): by analyzing the variations of a
product’s features, PCA can be used to identify the causes of product defects [122].
2. Machine monitoring (Know-why, machine level): by analyzing sensor data from
a machine, PCA can be used to detect incipient patterns in the data that indicate
potential issues with the machinery, such as wear and tear [123].
3. Process optimization (Know-why, process level): by analyzing variations in the pro-
cess data, PCA can be used to identify the most important factors that affect the pro-
cess, allowing the manufacturer to optimize the process and thereby reduce costs [124].
Autoencoder (AE) [119] is another popular method for reducing the dimensionality of
high-dimensional data. AE alone does not perform classification; instead, it provides a
compressed feature representation of high-dimensional data. The typical structure of AE
consists of an input layer, one hidden or encoding layer, one reconstruction or decoding
layer, and an output layer. The training strategy of AE includes encoding input data
into a latent representation that can reconstruct the input. To learn a compressed feature
representation of input data, AE tries to reduce the reconstruction error, that is, to minimize
the difference between the input and output data. An illustration of AE is shown in
Figure 11.

Figure 11. A pictorial representation of (a) an Autoencoder and (b) a Denoising Autoencoder. An
autoencoder is trained to reconstruct its input, while a denoising autoencoder is trained to reconstruct
a “clean” version of its input from a corrupted or “noisy” version of the input.

There are different types autoencoders that can be used for high-dimensional data. Stacked
Autoencoder (SAE) [119] is built by stacking multiple layers of AEs in such a way that the output
of one layer serves as the input of the subsequent layer. Denoising autoencoder (DAE) [125]
is a variant of AE that has a similar structure except for the input data. In DAE, the input
is corrupted by adding noise to it; however, the output is the original input signal without
noise. Therefore, unlike AE, DAE has the ability to recover the original input from a noisy
input signal. Convolutional autoencoder [126] is another interesting variant of AE, employing
convolutional layers to encode and decode high-dimensional data.
AEs can be used for a variety of applications in manufacturing, such as:
1. Anomaly detection (Know-what): an AE can be trained to reconstruct normal data
and detect abnormal data by measuring the reconstruction error, which allows the
manufacturer to detect and address issues such as product defects [124] and machinery
failure [127].
2. Feature selection (Know-why): an AE can be used to identify the most important
features in the data and remove the noise and irrelevant information, which can be
used for diagnosis of product defects or to detect events of interests [128].
3. Dimensionality reduction: an AE can be used to reduce the dimensionality of large
and complex datasets, making it easier to identify patterns and trends [129].
Appl. Sci. 2023, 13, 1903 15 of 32

Furthermore, AEs can be used in conjunction with other techniques, such as clustering
or classification, to improve the accuracy of prediction and enhance the interpretability of
the results [130]. Additionally, AEs can be used for data visualization. By reducing the
dimensionality of the data, AEs allow high-dimensional data to be visualized clearly and
interpretably [129] in a way that can be easily understood by non-technical stakeholders.
Clustering: The objective of clustering is to divide the set of datapoints into a number
of groups, ensuring that the datapoints within each group are similar to one another
and different from the datapoints in the other groups. Clustering methods are powerful
tools, allowing manufacturers examine large and complex datasets and gain meaningful
insights. There are different clustering methods available, each with their own strengths
and weaknesses, and the choice of method depends on the characteristics of the data
and the problem to be solved. Among the widely used clustering methods are Centroid-
based Clustering [120], Density-based Clustering [131], Distribution-based Clustering [132], and
Hierarchical Clustering [133]. Clustering algorithms have a wide range of applications in
manufacturing. For instance, clustering can be used to group manufactured inventory
parts according to different features [134] (Know-what). The obtained clusters can be used
as a guideline for warehouse space optimization [135]. Clustering can be used for anomaly
detection [136] (Know-what) and process optimization [137] (Know-how), and can be used
in conjunction with other techniques to improve the interpretability of results.
Association rule-based learning [121]: Association rule-based learning is an unsu-
pervised data-mining technique that finds important interactions among variables in a
dataset. It is capable of identifying hidden correlations in datasets by measuring degrees
of similarity. Hence, association rule-based learning is suitable in the Know-why stage in
manufacturing. For instance, association rule-based learning can be utilized to accurately
depict the relationship between quantifiable shop floor indicators and appropriate causes
of action under various conditions of machine utilization (Know-why, system level), which
can be used to establish an appropriate management strategy [138].

4.3. Semi-Supervised Learning Methods


Unsupervised learning methods do not have any input guidance during training,
which reduces labeling costs; however, their performance is normally less accurate. There-
fore, semi-supervised learning methods can be used to take advantage of the accuracy
achieved by supervised learning while limiting costs thanks to the reduction in labeling
effort. Therefore, researchers have turned to data augmentation [139,140] to enlarge dataset,
with the inputs and labels generated massively based on the existing dataset in a controlled
way while incurring no extra cost in the labeling phase. Taking an image with its label
as an example, it can be enriched by basic transformations such as rotation, translation,
flipping, noise injection, etc. It can be enriched by adversarial data augmentation, such
as by generating synthetic dataset using generative models, e.g., Generative Adversarial
Network (GAN) [141] and Variational AutoEncoder (VAE) [142], thereby obtaining new
images for training ML models at low cost. However, the improvements obtainable with
data augmentation are limited, and more real data are better than more synthetic data [143].
Therefore, increasing attention is being paid to the combination of supervised learning and
unsupervised learning, namely, semi-supervised learning, in which both unlabeled data
and labeled data are leveraged during training.
Semi-supervised learning methods can be generally divided into two groups: data
augmentation-based methods and semi-supervised mechanism-based methods. An overview
of semi-supervised methods is provided in Figure 12.
Data augmentation: through data augmentation, labeled data can be enlarged and aug-
mented by adding model predictions of newly unlabeled data with high confidence as pseudo-
labels, as shown in Figure 13. However, the model continues to be run in a fully supervised
manner. In addition, the quality of the pseudo-labels can highly affect model performance, and
incorrect pseudo-labels with high confidence are inevitable due to their nature. To improve the
quality of pseudo-labels, there are hybrid methods combining pseudo-labels and consistency
Appl. Sci. 2023, 13, 1903 16 of 32

regularization, such as MixMatch [144] and FixMatch [145]. Nevertheless, data augmentation-
based methods are simple, and there is no need to carefully design the loss. Therefore, data
augmentation-based methods can be potentially useful for non-experts in manufacturing for
enlarging labeled dataset when it is easy to collect massive amounts of unlabeled data.
Semi-supervised mechanisms: by contrast, semi-supervised mechanism-based meth-
ods are more focused on the mechanism of utilizing both labeled data and unlabeled
data. The principle of semi-supervised mechanisms is illustrated in Figure 14, where both
labeled data and unlabeled data can be model inputs while their losses are calculated in a
different way. Semi-supervised mechanism-based methods can be further categorized into
consistency-based methods, graph-based methods, and generative-based methods.

Figure 12. Overview of semi-supervised methods.

Figure 13. Data augmentation-based methods.

Figure 14. Semi-supervised mechanism-based methods.


Appl. Sci. 2023, 13, 1903 17 of 32

Consistency-based methods take advantage of the consistency of model outputs after


perturbations [146]; therefore, consistency regularization can be applied for unlabeled data.
Consistency constraint can be either imposed between the predictions from perturbed
inputs from the same sample, for instance, the π model [147], or between the predictions
from two models with the same architecture, such as MeanTeacher [148]. Thanks to the
perturbations in consistency-based methods, model generalization can be enhanced [149].
In terms of applications in manufacturing, depending on the output values consistency-
based methods can be used in the Know-what and Know-when stages. For instance,
consistency-based methods can be utilized in quality monitoring based on images (Know-
what, product level).
Graph-based methods aim to establish a graph from a dataset by denoting each data
point as a node, with the edge connecting two nodes representing the similarity between
them. Label propagation is then performed on the established graph, with the information
from labeled data used to infer the labels of the unlabeled data. Graph-based methods result
in the connected nodes being closer in the feature space, while disconnected nodes repel
each other. Therefore, graph-based methods can be used to address the problem of poor
class separation due to intra-class variations and inter-class similarities [18]. Consequently,
graph-based methods can be potentially useful for defect classification [18] (Know-what,
product level) or machine health state monitoring [150] (Know-what, machine level) where
there are problems with insufficient label information or poor class separation. However, it
should be noted that graph-based methods are normally transductive methods, meaning
that the constructed graph is only valid for the trained data and rebuilding the graph is
necessary when it comes to new data. Typical examples of graph-based methods include
Graph Neural Networks (GNNs) [151] and Graph Convolution Networks (GCNs) [152].
The main point of generative-based methods is to learn patterns from a dataset and to
model data distributions, allowing the model to be used to generate new samples. Then
during training, the model can be updated using the combination of the supervised loss
(for existing data with labels) and unsupervised loss (for synthetic data). An inherent
advantage of generative-based methods is that the labeled data can be enriched by a trained
model which has learned the data distribution. Therefore, generative-based methods are
well-suited for situations where it is difficult to collect labeled data, such as process fault
detection [153] (Know-what, process level) and anomaly detection in machinery [154]
(Know-what, machine level). Examples include the semi-supervised GAN series (SS-
GANs), such as Categorical Generative Adversarial Network (CatGAN) [155], Improved
GAN [156], and semi-supervised VAEs (SS-VAEs) [157].
Table A3 lists semi-supervised applications in manufacturing taken from the selected
documents in Section 2.2.

4.4. Reinforcement Learning Methods


Reinforcement Learning (RL) algorithms consist of two elements, namely, an agent
acting within an environment (see Figure 15). The agent is acting, and is therefore subject to
the desired learning process by directly interacting with and manipulating the environment.
Based on [158], the procedure of a learning cycle is as follows: first, the agent is presented
with an observation of the environment state st ∈ S; then, based on this observation (along
with internal decision making), the selection of an action at ∈ A. S refers to the state
space, that is, the set of possible observations that could occur in the environment. The
observation has to provide sufficient information on the current environment or system
state in order for the agent to select actions in an ideal way to solve the control problem. For
selecting the action, A refers to the action space, that is, the set of possible actions chosen by
the agent. After at is performed (in a given state st ), the environment moves to the resulting
state st+1 and the agent receives a reward rt+1 . Then, the reinforcement learning cycle
continues to iterate as shown in Figure 15. The agent aims to maximize the (discounted)
long-term cumulative reward by improving the selection of actions towards an optimum.
In other words, the RL agent wants to learn an optimal control policy for the environment.
Appl. Sci. 2023, 13, 1903 18 of 32

Figure 15. Overview of the Reinforcement Learning approach based on [158].

In general, RL approaches can be split into model-based, i.e., the agent has an internal
model of how the environment works, and model-free. The latter is most common thanks to
the advent of deep learning, and simplifies application, as feature selection can be applied.
Model-free approaches themselves can be divided into short value-based or policy-based
approaches by their approach to storing state-action value pairs, which are used to select
the action for optimal value return; the latter directly optimize the action selection policy.
In contrast to the other machine learning techniques, RL does not require large dataset,
only a clearly specified environment. Typically, an RL agent is trained on a simulation or
digital twin model [159]; after successful training, it can be implemented on the Know-how
level for its original purpose. Otherwise, the agent starts with random non-optimal actions,
leading to undesired system behavior.
Considering the aim of achieving the Know-how level for autonomous control in
processes, machines, or systems, RL is extremely important for applications in future
production. In addition, multi-agent RL is becoming of interest to the research commu-
nity [33], and can even be applied for controlling products [160]. However, RL remains
under-exploited in the industrial area, especially in respect to other machine learning
techniques [161].
As of now, applied approaches can be summarized as shown in Table A4. Note that
the applications reviewed here are implemented in a simulation or digital twin [159], and
features are manually crafted from raw data.

5. Challenges and Future Directions


A large number of ML use cases have shown the great potential for addressing complex
manufacturing problems, from knowing what is happening to knowing how employ self-
adapting or self-optimizing systems. The data-driven mechanisms in ML enable broader
applications in different fields as well as at different levels, from individual products to
whole systems. However, in spite of the great potential and advantages offered by ML and
numerous off-the-shelf ML models, there are critical challenges to overcome before the
successful application of ML in manufacturing can be realized. The following demonstrate
typical challenges that manufacturing industries might confront during the application
and deployment of ML-based solutions, along with corresponding future directions for
tackling these challenges (RQ3).
• Lack of data. Preparing the data used for ML is not a simple task, as the scale and the
quality of data can greatly affect the performance of ML models. The most common
challenge involves preparing a large amount of organized input data, and ensuring
high-quality labels if labels are needed. Despite manufacturing data becoming increas-
ingly more accessible due to the development of sensors and the Internet of Things,
gathering meaningful data is time-consuming and costly in many cases, for example,
fault detection and RUL prediction. This issue might be alleviated by the Synthetic
Minority Over-sampling Technique (SMOTE) [162]. However, SMOTE cannot capture
complex representative data, as it often relies on interpolation [163]. Data augmenta-
tion [139,164] or transfer learning [165] may address this problem. The aim of data
augmentation is to enlarge dataset by means of transforming data [139], by trans-
forming both data and labels, as with MixUp [166], or by generating synthetic data
using generative models [167,168]. On the contrary, instead of focusing on expanding
Appl. Sci. 2023, 13, 1903 19 of 32

data, transfer learning aims to leverage knowledge from similar external datasets. A
typically used method in transfer learning is parameter transfer, where a pretrained
model from a similar dataset is employed for initialization [165]. Another situation
involving lack of data is that certain data cannot be shared due to data privacy and
security issues. In confronting this problem, Federated Learning (FL) [169] might be a
potential opportunity to enable model training across multiple decentralized devices
while holding local data privately.
• Limited computing resources. The high performance of ML models always comes
with high computational complexity. In particular, obtaining high accuracy with a
neural network requires on millions or even billions of parameters [170]. However,
limited computing resources in industries makes it a challenge to deploy heavy ML
models in real-time industrial environments. Possible approaches include model
compression via pruning and sharing of model parameters [171] and knowledge
distillation [172]. Parameter pruning aims to reduce the number of model parameters
by removing redundant parameters without any effect on model performance. By
contrast, seeking the same goal, knowledge distillation focuses on distilling knowledge
from a cumbersome neural network to a lightweight network to allow it to be deployed
more easily with limited computing resources.
• Changing circumstances. Most ML applications in manufacturing focus only on model
development and verification in off-line environments. However, when deploying these
models in running production, their performance may be degraded due to changing cir-
cumstances, leading to changes in data distribution, that is, drift [173,174]. Therefore, man-
ual model adjustment over time, which is time-consuming, is usually unavoidable [175].
However, this could be addressed in the future by automatic model adaption [174], in
which data drifts are automatically detected and handled with less resources.
• Interpretability of results. Many expectations have been placed on ML to overcome all
types of problems without the need for prior knowledge. In particular, ML models are
expected to directly learn higher level knowledge such as Know-when and Know-how,
which is difficult for human beings to obtain in manufacturing. However, without
the foundations of early-stage knowledge and an understanding of the data, the
results inferred from big data by black-box ML models are meaningless and unreliable.
For instance, predictions blindly obtained from all data, including both relevant
and irrelevant data, might even degrade performance due to the GIGO (garbage in,
garbage out) phenomenon [176]. To overcome this problem, future directions within
ML development might include incorporating physical models into ML models [177]
or obtaining Four-know knowledge successively.
• Uncertainty of results. Related to the challenge of interpretability is the challenge of
uncertain results. The success of manufacturing depends heavily on the quality of
the resulting products. As every manufacturing process has a degree of variability,
almost all industrial manufacturers use statistical process control (SPC) to ensure a
stable and defined quality of products [178]. A central element of statistical process
control is the determination and handling of statistical uncertainty. The uncertainty
of ML results often cannot be quantified reliably and efficiently, even with today’s
state-of-the-art [179–181]. Furthermore, model complexity and severe non-linearity
in ML can hinder the evaluation of uncertainty [182]. Although there are promising
approaches, e.g., Gaussian mixture models for NN [183,184] and Probabilistic Neural
Network (PNN) [184], or the use of Baysian Networks [180], there are several limita-
tions limiting potential applications, such as high computational cost and simplified
assumptions [184]. Therefore, future research needs to make progress on the general
theory of integrating uncertainty into ML methods to allow manufacturing in order to
ensure high quality and stability in production.
To summarize, while ML is a fairly open tool which can be used to handle a variety
of problems in manufacturing, it is necessary to have an understanding of the hidden
challenges in ML application in order to provide more realistic and robust outcomes. For in-
Appl. Sci. 2023, 13, 1903 20 of 32

stance, early in ML application in manufacturing, one might face the problem of lacking data.
During the deployment of ML-based solutions, one might confront challenges around inte-
grating the solution into the industrial environment. After deployment, one might encounter
the challenge of evaluating ML results on product and process in terms of interpretability
and uncertainty. The future directions pointed out in this review can help to address the
above-mentioned challenges and ensure reliable improvements in manufacturing contexts.

6. Conclusions
It is fully recognized that ML is playing an increasingly critical role in the digitization
of manufacturing industries towards Industry 4.0, leading to improved quality, produc-
tivity, and efficiency. This review has paper aimed to address the issue of how ML can
improve manufacturing, posing three research questions related to the above issue in
the introduction. To address these research questions, we carried out a literature review
assessing the state-of-the-art based on 1348 published scientific articles.
To answer RQ1, we first introduced the concepts of the ‘Four-Know’ (Know-what,
Know-why, Know-when, Know-how) and ‘Four-Level’ (Product, Process, Machine, System)
categories to help formulate ML tasks in manufacturing. By mapping ML use cases into the
Four-Know and Four-Level matrix, we provide an understanding of typical ML use cases
and their potential benefits for improving manufacturing. To further support RQ1, the
identified ML studies were classified using the ’Four-Know’ and ’Four-Level’ perspective
to provide an overview of ML publications in manufacturing. The results showed that
current ML applications are mainly focused on the product level, in particular in terms
of Know-what and Know-when. In addition, considerable growth in Know-how was
observed at the process and system levels, which might be correlated to higher demand for
adaptability to changes on these levels.
To fill the gap between academic research and manufacturing industries, we provided
an actionable pipeline for the implementation of ML solutions by production engineers from
ideation through to deployment, thereby answering RQ2. To further explain the ’model
training’ step, which is the core stage in the pipeline, a holistic review of ML methods
was provided, including supervised, semi-supervised, unsupervised, and reinforcement
learning methods along with their typical applications in manufacturing. We hope that this
can provide support in method selection for decision-makers considering ML solutions.
Finally, to answer RQ3, we uncovered the current challenges that manufacturing
industry is likely to encounter during application and deployment, and provided possible
future directions for tackling these challenges as possible developments for ensuring more
reliable and robust outcomes in manufacturing.

Author Contributions: Conceptualization, T.C., O.J.J., M.C.M., V.S. and G.F.; methodology, T.C. and
S.S.; formal analysis, T.C. and S.S.; writing—original draft preparation, T.C., V.S., S.S., M.C.M., O.J.J.
and F.S.; writing—review and editing, T.C., V.S., M.C.M., S.S., O.J.J., M.C., G.F., G.T., J.J.A.M. and F.S.;
supervision, M.C., G.F., G.T., J.J.A.M. and F.S.; funding acquisition, G.T. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was funded by a European Training Network supported by Horizon 2020,
grant number 814225.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: This research work was undertaken in the context of the DIGIMAN4.0 project
(“DIGItal MANufacturing Technologies for Zero-defect Industry 4.0 Production”, https://www.
digiman4-0.mek.dtu.dk/, accessed on 1 January 2023). DIGIMAN4.0 is a European Training Network
supported by Horizon 2020, the EU Framework Programme for Research and Innovation (Project
ID: 814225).
Conflicts of Interest: The authors declare no conflict of interest.
Appl. Sci. 2023, 13, 1903 21 of 32

Appendix A

Table A1. Categories of supervised learning applications.

Know- Know- Know- Know-


Ref. Year Level Data Type Method Type Case Field
What Why When How
[104] 2018 Product X Image Kernel Defect detection Metallic powder bed fusion
Metal frame process in mobile device
[185] 2018 Product X Tabular Kernel Product monitoring
manufacturing
Temperature prediction and po-
[186] 2020 Process X X X Time series, Image Kernel Additive manufacturing
tential anomaly detection
Product,
[19] 2022 X X Tabular Kernel Fault detection and classification Semiconductor Etch Equipment
Process
[105] 2022 Product X Tabular Kernel Quality prediction Additive manufacturing
[106] 2022 Machine X Time series Kernel Wear prediction Metal forming
[187] 2022 System X X Time series Kernel Content prediction Steel making
[114] 2018 Machine X X Time series (Image) NN Fault diagnosis Motor bearning and pump
[188] 2020 System X X Image NN Cost estimation
[112] 2020 Product X Image NN Defect detection Battery manufacturing
[189] 2022 Product X Time series NN Quality assurance Fused deposition modeling
[190] 2022 Process X Time series NN Process optimization Wire arc additive manufacturing
[191] 2022 Process X X Tabular NN Parameter optimization Laser powder bed fusion
[192] 2022 Process X Image NN Object detection Robotic grasp
[193] 2022 Product X Image NN, kernel Defect detection Roller manufacturing
[194] 2022 Machine X Time series (Image) NN, kernel Tool condition monitoring Machining
[195] 2019 Product X Time series Tree Material removal prediction Robotic grinding
[196] 2022 Product X Image (Tabular) Tree Porosity prediction Powder-bed additive manufacturing
[197] 2019 Product X Image Probabilistic Online quality inspection Powder-bed additive manufacturing
[198] 2018 System X Tabular Hybrid Scheduling Flexible Manufacturing Systems (FMSs)
Appl. Sci. 2023, 13, 1903 22 of 32

Table A2. Categories of unsupervised learning applications.

Know- Know- Know- Know-


Ref. Year Level Data Type Method Type Case Field
What Why When How
[120] 2021 Machine X Time series Clustering Tool Condition clustering Autonomous manufacturing
[131] 2021 Machine X Time series Clustering Tool health monitoring Machine tool health Monitoring
[132] 2019 Machine X X Time series Clustering Defect Identification Manufacturing systems
[199] 2021 Process X Tabular, Time series Clustering Condition monitoring Manufacturing Condition monitoring
[133] 2020 System X Time series Clustering Condition monitoring Manufacturing Condition monitoring
[125] 2018 Product X Image, Text Autoencoder Defect Identification Fabric industry
[119] 2019 Product X Image Autoencoder Defect Identification Automatic Optical Inspection
[200] 2021 Product X Image Autoencoder Defect Identification Printed circuit board manufacturing
[201] 2022 Machine X X Tabular, Time series Autoencoder Anomaly detection Steel rolling Process
[126] 2022 Process X X Image Autoencoder Anomaly detection Industrial Anomaly detection
[126] 2022 Process X X Image, Text Autoencoder Anomaly detection Semi conductor manufacturing
[202] 2022 Machine X Time series PCA Predictive maintenance Fan-motor system
[118] 2022 Machine X X Time series PCA Anomaly detection Programmable logic controllers
[121] 2015 Process X Tabular Association rule Predictive maintenance Wooden door manufacturing
Appl. Sci. 2023, 13, 1903 23 of 32

Table A3. Categories of semi-supervised learning applications.

Know- Know- Know- Know-


Ref. Year Level Data Type Method Type Case Field
What Why When How
[203] 2020 Product X Image Data augmentation Quality control Automated Surface Inspection
[204] 2021 Process X Image Data augmentation Measurement in process Positioning of welding seams
[205] 2019 System X Tabular Data augmentation Energy consumption modelling Steel industry
Product,
[206] 2020 X Time series Data augmentation Quality prediction Continuous-flow manufacturing.
System
[207] 2021 Machine X Time series Consistency-based Predictive quality control Semiconductor manufacturing
[149] 2020 Product X Image Consistency-based Quality monitoring Metal additive manufacturing
[18] 2021 Product X Image Graph-based Quality control Automated Surface Inspection
[150] 2022 Machine X X Time series Graph-based Machine health state diagnosis Manipulator
[208] 2022 Machine X X Tabular Graph-based Predict tool tip dynamics Machine tool
Assessing manufacturability of
[209] 2021 Product X Image Generative-based Direct metal laser sintering process
cellular structures
[210] 2019 Product X Time series Generative-based Quality inferred from process laser powder-bed fusion
[211] 2020 Product X Image Generative-based Quality diagnosis Wafer fabrication
[212] 2021 Product X Image Generative-based Quality control Automated Surface Inspection
[213] 2020 Machine X Time series Generative-based Remaining useful life prognostics Turbofan engine and rolling bearing
Vacuum system in styrene petrochemical
[214] 2021 Machine X X Tabular Generative-based Machine condition monitoring
plant
Anomaly detection for predictive
[153] 2021 Machine X X Time series Generative-based Press machine
maintenance
[154] 2022 Process X Time series (image) Generative-based Process fault detection Die casting process
Appl. Sci. 2023, 13, 1903 24 of 32

Table A4. Categories of reinforcement learning applications.

Know- Know- Know- Know-


Ref. Year Level Data Type Method Type Case Field
What Why When How
[32] 2021 Process X X Tabular Value-based Quality control Statistical Process Control
[215] 2022 System X X Tabular Value-based Scheduling Semiconductor fab
[216] 2021 System X Tabular Value-based Throughput control Flow shop
[217] 2021 Machine X Tabular Value-based Scheduling & Maintenance Multi-state single machine
[34] 2020 System X Tabular Value-based Quality Control & Maintenance Production system
[218] 2022 System X Tabular Value-based Lead time management Flow shop
[219] 2020 Process X Tabular Value-based Robotic arm control Soft fabric manufacturing
[220] 2021 System X X Tabular Value-based Layout planning Greenfield factories
[221] 2020 Machine X Tabular Value-based Maintenance scheduling Preventive maintenance
[33] 2022 Machine X X Tabular Policy-based Maintenance scheduling Parallel machines
[222] 2021 Process X Tabular Policy-based Improving efficiency Automated product disassembly
[223] 2021 System X Tabular Policy-based Dispatching Job shop
[224] 2022 System X Tabular Policy-based Scheduling & maintenance Semiconductor fab
[225] 2022 System X Tabular Policy-based Yield optimization Multi-agent RL
[57] 2022 System X Tabular Policy-based Human Worker Control Flow shop
[59] 2022 System X Tabular Policy-based Scheduling & dispatching Disassembly job shop
[160] 2021 Product X Tabular Policy-based Multi-agent production control Job shop
[226] 2022 Process X Tabular Both Parameter optimisation Manufacturing processes
[227] 2019 Process X Tabular Both Online parameter optimisation Injection molding
[228] 2022 System X Tabular Both scheduling Matrix production system
Appl. Sci. 2023, 13, 1903 25 of 32

References
1. Abele, E.; Reinhart, G. Zukunft der Produktion: Herausforderungen, Forschungsfelder, Chancen; Hanser: München, Germany, 2011.
2. Zizic, M.C.; Mladineo, M.; Gjeldum, N.; Celent, L. From industry 4.0 towards industry 5.0: A review and analysis of paradigm
shift for the people, organization and technology. Energies 2022, 15, 5221. [CrossRef]
3. Huang, S.; Wang, B.; Li, X.; Zheng, P.; Mourtzis, D.; Wang, L. Industry 5.0 and Society 5.0—Comparison, complementation and
co-evolution. J. Manuf. Syst. 2022, 64, 424–428. [CrossRef]
4. Vukovic, M.; Mazzei, D.; Chessa, S.; Fantoni, G. Digital Twins in Industrial IoT: A survey of the state of the art and of relevant
standards. In Proceedings of the 2021 IEEE International Conference on Communications Workshops (ICC Workshops), Montreal,
QC, Canada, 14–23 June 2021. [CrossRef]
5. Mourtzis, D.; Fotia, S.; Boli, N.; Vlachou, E. Modelling and quantification of industry 4.0 manufacturing complexity based on
information theory: A robotics case study. Int. J. Prod. Res. 2019, 57, 6908–6921. [CrossRef]
6. Galin, R.; Meshcheryakov, R.; Kamesheva, S.; Samoshina, A. Cobots and the benefits of their implementation in intelligent
manufacturing. IOP Conf. Ser. Mater. Sci. Eng. 2020, 862, 032075. [CrossRef]
7. May, M.C.; Schmidt, S.; Kuhnle, A.; Stricker, N.; Lanza, G. Product Generation Module: Automated Production Planning for
optimized workload and increased efficiency in Matrix Production Systems. Procedia CIRP 2020, 96, 45–50. [CrossRef]
8. Lu, Y. Industry 4.0: A survey on technologies, applications and open research issues. J. Ind. Inf. Integr. 2017, 6, 1–10. [CrossRef]
9. Miqueo, A.; Torralba, M.; Yagüe-Fabra, J.A. Lean manual assembly 4.0: A systematic review. Appl. Sci. 2020, 10, 8555. [CrossRef]
10. Wuest, T.; Weimer, D.; Irgens, C.; Thoben, K.D. Machine learning in manufacturing: Advantages, challenges, and applications.
Prod. Manuf. Res. 2016, 4, 23–45. [CrossRef]
11. Rai, R.; Tiwari, M.K.; Ivanov, D.; Dolgui, A. Machine learning in manufacturing and industry 4.0 applications. Int. J. Prod. Res.
2021, 59, 4773–4778. [CrossRef]
12. Bertolini, M.; Mezzogori, D.; Neroni, M.; Zammori, F. Machine Learning for industrial applications: A comprehensive literature
review. Expert Syst. Appl. 2021, 175, 114820. [CrossRef]
13. Wang, J.; Ma, Y.; Zhang, L.; Gao, R.X.; Wu, D. Deep learning for smart manufacturing : Methods and applications. J. Manuf. Syst.
2018, 48, 144–156. [CrossRef]
14. Dogan, A.; Birant, D. Machine learning and data mining in manufacturing. Expert Syst. Appl. 2021, 166, 114060. [CrossRef]
15. Alshangiti, M.; Sapkota, H.; Murukannaiah, P.K.; Liu, X.; Yu, Q. Why is developing machine learning applications challenging?
a study on stack overflow posts. In Proceedings of the 2019 ACM/IEEE International Symposium on Empirical Software
Engineering and Measurement (ESEM), Porto de Galinhas, Brazil, 19–20 September 2019; pp. 1–11.
16. Zeller, V.; Hocken, C.; Stich, V. Acatech Industrie 4.0 maturity index—A multidimensional maturity model. In Proceedings of the
IFIP International Conference on Advances in Production Management Systems, Seoul, Republic of Korea, 26–30 August 2018;
Springer: Cham, Switzerland, 2018; pp. 105–113.
17. Yang, L.; Fan, J.; Huo, B.; Li, E.; Liu, Y. A nondestructive automatic defect detection method with pixelwise segmentation.
Knowl.-Based Syst. 2022, 242, 108338. [CrossRef]
18. Wang, Y.; Gao, L.; Gao, Y.; Li, X. A new graph-based semi-supervised method for surface defect classification. Robot. Comput.
Integr. Manuf. 2021, 68, 102083. [CrossRef]
19. Kim, S.H.; Kim, C.Y.; Seol, D.H.; Choi, J.E.; Hong, S.J. Machine Learning-Based Process-Level Fault Detection and Part-Level
Fault Classification in Semiconductor Etch Equipment. IEEE Trans. Semicond. Manuf. 2022, 35, 174–185. [CrossRef]
20. Peng, S.; Feng, Q.M. Reinforcement learning with Gaussian processes for condition-based maintenance. Comput. Ind. Eng. 2021,
158, 107321. [CrossRef]
21. Zheng, W.; Liu, Y.; Gao, Z.; Yang, J. Just-in-time semi-supervised soft sensor for quality prediction in industrial rubber mixers.
Chemom. Intell. Lab. Syst. 2018, 180, 36–41. [CrossRef]
22. Kang, P.; Kim, D.; Cho, S. Semi-supervised support vector regression based on self-training with label uncertainty: An application to
virtual metrology in semiconductor manufacturing. Expert Syst. Appl. 2016, 51, 85–106. [CrossRef]
23. Srivastava, A.K.; Patra, P.K.; Jha, R. AHSS applications in Industry 4.0: Determination of optimum processing parameters during
coiling process through unsupervised machine learning approach. Mater. Today Commun. 2022, 31, 103625. [CrossRef]
24. Antomarioni, S.; Ciarapica, F.E.; Bevilacqua, M. Association rules and social network analysis for supporting failure mode effects
and criticality analysis : Framework development and insights from an onshore platform. Saf. Sci. 2022, 150, 105711. [CrossRef]
25. Pan, R.; Li, X.; Chakrabarty, K. Semi-Supervised Root-Cause Analysis with Co-Training for Integrated Systems. In Proceedings of
the 2022 IEEE 40th VLSI Test Symposium (VTS), San Diego, CA, USA, 25–27 April 2022. [CrossRef]
26. Chen, R.; Lu, Y.; Witherell, P.; Simpson, T.W.; Kumara, S.; Yang, H. Ontology-Driven Learning of Bayesian Network for Causal
Inference and Quality Assurance in Additive Manufacturing. IEEE Robot. Autom. Lett. 2021, 6, 6032–6038. [CrossRef]
27. Sikder, S.; Mukherjee, I.; Panja, S.C. A synergistic Mahalanobis–Taguchi system and support vector regression based predictive
multivariate manufacturing process quality control approach. J. Manuf. Syst. 2020, 57, 323–337. [CrossRef]
28. Cerquitelli, T.; Ventura, F.; Apiletti, D.; Baralis, E.; Macii, E.; Poncino, M. Enhancing manufacturing intelligence through an
unsupervised data-driven methodology for cyclic industrial processes. Expert Syst. Appl. 2021, 182, 115269. [CrossRef]
29. Kolokas, N.; Vafeiadis, T.; Ioannidis, D.; Tzovaras, D. A generic fault prognostics algorithm for manufacturing industries using
unsupervised machine learning classifiers. Simul. Model. Pract. Theory 2020, 103, 102109. [CrossRef]
Appl. Sci. 2023, 13, 1903 26 of 32

30. Verstraete, D.; Droguett, E.; Modarres, M. A deep adversarial approach based on multisensor fusion for remaining useful life
prognostics. In Proceedings of the 29th European Safety and Reliability Conference (ESREL 2019), Hannover, Germany, 22–26
September 2020; pp. 1072–1077. [CrossRef]
31. Wu, D.; Jennings, C.; Terpenny, J.; Gao, R.X.; Kumara, S. A Comparative Study on Machine Learning Algorithms for Smart
Manufacturing: Tool Wear Prediction Using Random Forests. J. Manuf. Sci. Eng. Trans. ASME 2017, 139, 071018. [CrossRef]
32. Viharos, Z.J.; Jakab, R. Reinforcement Learning for Statistical Process Control in Manufacturing. Meas. J. Int. Meas. Confed. 2021,
182, 109616. [CrossRef]
33. Luis, M.; Rodríguez, R.; Kubler, S.; Giorgio, A.D.; Cordy, M.; Robert, J.; Le, Y. Multi-agent deep reinforcement learning based
Predictive Maintenance on parallel machines. Robot. Comput. Integr. Manuf. 2022, 78, 102406.
34. Paraschos, P.D.; Koulinas, G.K.; Koulouriotis, D.E. Reinforcement learning for combined production-maintenance and quality
control of a manufacturing system with deterioration failures. J. Manuf. Syst. 2020, 56, 470–483. [CrossRef]
35. Liu, Y.H.; Huang, H.P.; Lin, Y.S. Dynamic scheduling of flexible manufacturing system using support vector machines. In
Proceedings of the 2005 IEEE Conference on Automation Science and Engineering, IEEE-CASE 2005, Edmonton, AB, Canada, 1–2
August 2005; Volume 2005, pp. 387–392. [CrossRef]
36. Zhou, G.; Chen, Z.; Zhang, C.; Chang, F. An adaptive ensemble deep forest based dynamic scheduling strategy for low carbon
flexible job shop under recessive disturbance. J. Clean. Prod. 2022, 337, 130541. [CrossRef]
37. de la Rosa, F.L.; Gómez-Sirvent, J.L.; Sánchez-Reolid, R.; Morales, R.; Fernández-Caballero, A. Geometric transformation-based
data augmentation on defect classification of segmented images of semiconductor materials using a ResNet50 convolutional
neural network. Expert Syst. Appl. 2022, 206, 117731. [CrossRef]
38. Krahe, C.; Marinov, M.; Schmutz, T.; Hermann, Y.; Bonny, M.; May, M.; Lanza, G. AI based geometric similarity search supporting
component reuse in engineering design. Procedia CIRP 2022, 109, 275–280. [CrossRef]
39. Onler, R.; Koca, A.S.; Kirim, B.; Soylemez, E. Multi-objective optimization of binder jet additive manufacturing of Co-Cr-Mo
using machine learning. Int. J. Adv. Manuf. Technol. 2022, 119, 1091–1108. [CrossRef]
40. Jadidi, A.; Mi, Y.; Sikström, F.; Nilsen, M.; Ancona, A. Beam Offset Detection in Laser Stake Welding of Tee Joints Using Machine
Learning and Spectrometer Measurements. Sensors 2022, 22, 3881. [CrossRef]
41. Sanchez, S.; Rengasamy, D.; Hyde, C.J.; Figueredo, G.P.; Rothwell, B. Machine learning to determine the main factors affecting
creep rates in laser powder bed fusion. J. Intell. Manuf. 2021, 32, 2353–2373. [CrossRef]
42. Verma, S.; Misra, J.P.; Popli, D. Modeling of friction stir welding of aviation grade aluminium alloy using machine learning
approaches. Int. J. Model. Simul. 2022, 42, 1–8. [CrossRef]
43. Gerling, A.; Ziekow, H.; Hess, A.; Schreier, U.; Seiffer, C.; Abdeslam, D.O. Comparison of algorithms for error prediction in
manufacturing with automl and a cost-based metric. J. Intell. Manuf. 2022, 33, 555–573. [CrossRef]
44. Akbari, P.; Ogoke, F.; Kao, N.Y.; Meidani, K.; Yeh, C.Y.; Lee, W.; Farimani, A.B. MeltpoolNet: Melt pool characteristic prediction in
Metal Additive Manufacturing using machine learning. Addit. Manuf. 2022, 55, 102817. [CrossRef]
45. Dittrich, M.A.; Uhlich, F.; Denkena, B. Self-optimizing tool path generation for 5-axis machining processes. CIRP J. Manuf. Sci.
Technol. 2019, 24, 49–54. [CrossRef]
46. Xi, Z. Model predictive control of melt pool size for the laser powder bed fusion process under process uncertainty. ASCE-ASME
J. Risk Uncertain. Eng. Syst. Part B Mech. Eng. 2022, 8, 011103. [CrossRef]
47. Li, X.; Liu, X.; Yue, C.; Liu, S.; Zhang, B.; Li, R.; Liang, S.Y.; Wang, L. A data-driven approach for tool wear recognition and
quantitative prediction based on radar map feature fusion. Measurement 2021, 185, 110072. [CrossRef]
48. Xia, B.; Wang, K.; Xu, A.; Zeng, P.; Yang, N.; Li, B. Intelligent Fault Diagnosis for Bearings of Industrial Robot Joints Under Varying
Working Conditions Based on Deep Adversarial Domain Adaptation. IEEE Trans. Instrum. Meas. 2022, 71, 1–13. [CrossRef]
49. May, M.C.; Neidhöfer, J.; Körner, T.; Schäfer, L.; Lanza, G. Applying Natural Language Processing in Manufacturing. Procedia
CIRP 2022, 115, 184–189. [CrossRef]
50. Xu, X.; Li, X.; Ming, W.; Chen, M. A novel multi-scale CNN and attention mechanism method with multi-sensor signal for
remaining useful life prediction. Comput. Ind. Eng. 2022, 169, 108204. [CrossRef]
51. Shah, M.; Vakharia, V.; Chaudhari, R.; Vora, J.; Pimenov, D.Y.; Giasin, K. Tool wear prediction in face milling of stainless steel
using singular generative adversarial network and LSTM deep learning models. Int. J. Adv. Manuf. Technol. 2022, 121, 723–736.
[CrossRef]
52. Verl, A.; Steinle, L. Adaptive compensation of the transmission errors in rack-and-pinion drives. CIRP Ann. 2022, 71, 345–348.
[CrossRef]
53. Frigerio, N.; Cornaggia, C.F.; Matta, A. An adaptive policy for on-line Energy-Efficient Control of machine tools under throughput
constraint. J. Clean. Prod. 2021, 287, 125367. [CrossRef]
54. Bozcan, I.; Korndorfer, C.; Madsen, M.W.; Kayacan, E. Score-Based Anomaly Detection for Smart Manufacturing Systems.
IEEE/ASME Trans. Mechatron. 2022, 27, 5233–5242. [CrossRef]
55. Bokrantz, J.; Skoogh, A.; Nawcki, M.; Ito, A.; Hagstr, M.; Gandhi, K.; Bergsj, D. Improved root cause analysis supporting resilient
production systems. J. Manuf. Syst. 2022, 64, 468–478. [CrossRef]
56. Long, T.; Li, Y.; Chen, J. Productivity prediction in aircraft final assembly lines: Comparisons and insights in different productivity
ranges. J. Manuf. Syst. 2022, 62, 377–389. [CrossRef]
Appl. Sci. 2023, 13, 1903 27 of 32

57. Overbeck, L.; Hugues, A.; May, M.C.; Kuhnle, A.; Lanza, G. Reinforcement Learning Based Production Control of Semi-automated
Manufacturing Systems. Procedia CIRP 2021, 103, 170–175. [CrossRef]
58. May, M.C.; Behnen, L.; Holzer, A.; Kuhnle, A.; Lanza, G. Multi-variate time-series for time constraint adherence prediction in
complex job shops. Procedia CIRP 2021, 103, 55–60. [CrossRef]
59. Wurster, M.; Michel, M.; May, M.C.; Kuhnle, A.; Stricker, N.; Lanza, G. Modelling and condition-based control of a flexible and
hybrid disassembly system with manual and autonomous workstations using reinforcement learning. J. Intell. Manuf. 2022,
33, 575–591. [CrossRef]
60. Liberati, A.; Altman, D.G.; Tetzlaff, J.; Mulrow, C.; Gøtzsche, P.C.; Ioannidis, J.P.; Clarke, M.; Devereaux, P.J.; Kleijnen, J.; Moher, D.
The PRISMA statement for reporting systematic reviews and meta-analyses of studies that evaluate health care interventions:
Explanation and elaboration. J. Clin. Epidemiol. 2009, 62, e1–e34. [CrossRef]
61. Sampath, V.; Maurtua, I.; Aguilar Martín, J.J.; Gutierrez, A. A survey on generative adversarial networks for imbalance problems
in computer vision tasks. J. Big Data 2021, 8, 27. [CrossRef]
62. Polyzotis, N.; Roy, S.; Whang, S.E.; Zinkevich, M. Data lifecycle challenges in production machine learning: A survey. ACM
Sigmod Rec. 2018, 47, 17–28. [CrossRef]
63. Wang, Z.; Oates, T. Imaging time-series to improve classification and imputation. In Proceedings of the Twenty-Fourth
International Joint Conference on Artificial Intelligence, Buenos Aires, Argentina, 25–31 July 2015.
64. Lee, G.; Flowers, M.; Dyer, M. Learning distributed representations of conceptual knowledge. In Proceedings of the International
1989 Joint Conference on Neural Networks, Washington, DC, USA, 18–22 June 1989. [CrossRef]
65. Zhu, Y.; Brettin, T.; Xia, F.; Partin, A.; Shukla, M.; Yoo, H.; Evrard, Y.A.; Doroshow, J.H.; Stevens, R.L. Converting tabular data into
images for deep learning with convolutional neural networks. Sci. Rep. 2021, 11, 11325. [CrossRef]
66. Sharma, A.; Vans, E.; Shigemizu, D.; Boroevich, K.A.; Tsunoda, T. DeepInsight: A methodology to transform a non-image data to
an image for convolution neural network architecture. Sci. Rep. 2019, 9, 11399. [CrossRef]
67. LeCun, Y.; Bottou, L.; Bengio, Y.; Haffner, P. Gradient-based learning applied to document recognition. Proc. IEEE 1998,
86, 2278–2324. [CrossRef]
68. Nanni, L.; Ghidoni, S.; Brahnam, S. Handcrafted vs. non-handcrafted features for computer vision classification. Pattern Recognit.
2017, 71, 158–172. [CrossRef]
69. Alkinani, M.H.; Khan, W.Z.; Arshad, Q.; Raza, M. HSDDD: A Hybrid Scheme for the Detection of Distracted Driving through
Fusion of Deep Learning and Handcrafted Features. Sensors 2022, 22, 1864. [CrossRef]
70. Chen, Z.; Zhang, L.; Cao, Z.; Guo, J. Distilling the Knowledge from Handcrafted Features for Human Activity Recognition. IEEE
Trans. Ind. Inform. 2018, 14, 4334–4342. [CrossRef]
71. Albawi, S.; Mohammed, T.A.; Al-Zawi, S. Understanding of a convolutional neural network. In Proceedings of the 2017
International Conference on Engineering and Technology (ICET), Antalya, Turkey, 21–23 August 2017; pp. 1–6.
72. Pearson, K. LIII. On lines and planes of closest fit to systems of points in space. Lond. Edinb. Dublin Philos. Mag. J. Sci. 1901,
2, 559–572. [CrossRef]
73. Comon, P. Independent component analysis, a new concept? Signal Process. 1994, 36, 287–314. [CrossRef]
74. LeCun, Y.; Boser, B.; Denker, J.S.; Henderson, D.; Howard, R.E.; Hubbard, W.; Jackel, L.D. Backpropagation applied to handwritten
zip code recognition. Neural Comput. 1989, 1, 541–551. [CrossRef]
75. Mikolov, T.; Karafiát, M.; Burget, L.; Cernockỳ, J.; Khudanpur, S. Recurrent neural network based language model. In Interspeech;
Makuhari: Chiba-city, Japan, 2010; Volume 2, pp. 1045–1048.
76. Ojala, T.; Pietikainen, M.; Maenpaa, T. Multiresolution gray-scale and rotation invariant texture classification with local binary
patterns. IEEE Trans. Pattern Anal. Mach. Intell. 2002, 24, 971–987. [CrossRef]
77. Lowe, D.G. Distinctive image features from scale-invariant keypoints. Int. J. Comput. Vis. 2004, 60, 91–110. [CrossRef]
78. Dalal, N.; Triggs, B. Histograms of oriented gradients for human detection. In Proceedings of the 2005 IEEE Computer Society
Conference on Computer Vision and Pattern Recognition (CVPR’05), San Diego, CA, USA, 20–26 June 2005; Volume 1, pp. 886–893.
79. Shensa, M.J. The discrete wavelet transform: Wedding the a trous and Mallat algorithms. IEEE Trans. Signal Process. 1992,
40, 2464–2482. [CrossRef]
80. Gröchenig, K. The short-time Fourier transform. In Foundations of Time-Frequency Analysis; Springer Science & Business Media:
Berlin/Heidelberg, Germany, 2001; pp. 37–58.
81. Harris, Z.S. Distributional structure. Word 1954, 10, 146–162. [CrossRef]
82. Mikolov, T.; Chen, K.; Corrado, G.; Dean, J. Efficient estimation of word representations in vector space. arXiv 2013,
arXiv:1301.3781.
83. Liu, D.; Kong, H.; Luo, X.; Liu, W.; Subramaniam, R. Bringing AI to edge: From deep learning’s perspective. Neurocomputing
2021, 485, 297–320. [CrossRef]
84. Gray, R.M.; Neuhoff, D.L. Quantization. IEEE Trans. Inf. Theory 1998, 44, 2325–2383. [CrossRef]
85. Sampath, V.; Maurtua, I.; Aguilar Martín, J.J.; Iriondo, A.; Lluvia, I.; Rivera, A. Vision Transformer based knowledge distillation for
fasteners defect detection. In Proceedings of the 2022 International Conference on Electrical, Computer and Energy Technologies
(ICECET), Prague, Czech Republic, 20–22 July 2022; pp. 1–6.
86. Shelden, R. Decision Tree. Chem. Eng. Prog. 1970, 66, 8.
87. Breiman, L. Bagging predictors. Mach. Learn. 1996, 24, 123–140. [CrossRef]
Appl. Sci. 2023, 13, 1903 28 of 32

88. Freund, Y.; Schapire, R.E.; others. Experiments with a new boosting algorithm.. In Proceedings of the Thirteenth International
Conference on International Conference on Machine Learning (ICML’96), Bari, Italy, 3–6 July 1996; Volume 96, pp. 148–156.
89. Ho, T.K. Random decision forests. In Proceedings of the 3rd International Conference on Document Analysis and Recognition,
Montreal, QC, Canada, 14–16 August 1995; Volume 1, pp. 278–282.
90. Chen, T.; Guestrin, C. XGBoost. In Proceedings of the 22nd ACM SIGKDD International Conference on Knowledge Discovery
and Data Mining, San Francisco, CA, USA, 13–17 August 2016. [CrossRef]
91. Choi, S.; Battulga, L.; Nasridinov, A.; Yoo, K.H. A decision tree approach for identifying defective products in the manufacturing
process. Int. J. Contents 2017, 13, 57–65.
92. Sugumaran, V.; Muralidharan, V.; Ramachandran, K. Feature selection using decision tree and classification through proximal
support vector machine for fault diagnostics of roller bearing. Mech. Syst. Signal Process. 2007, 21, 930–942. [CrossRef]
93. Hung, Y.H. Improved ensemble-learning algorithm for predictive maintenance in the manufacturing process. Appl. Sci. 2021,
11, 6832. [CrossRef]
94. Močkus, J. On bayesian methods for seeking the extremum. In Optimization Techniques IFIP Technical Conference Novosibirsk,
Novosibirsk, Russia, 1–7 July 1974; Marchuk, G.I., Ed.; Springer: Berlin/Heidelberg, Germany, 1975; pp. 400–404.
95. Baum, L.E.; Petrie, T. Statistical inference for probabilistic functions of finite state Markov chains. Ann. Math. Stat. 1966,
37, 1554–1563. [CrossRef]
96. Papananias, M.; McLeay, T.E.; Mahfouf, M.; Kadirkamanathan, V. A Bayesian framework to estimate part quality and associated
uncertainties in multistage manufacturing. Comput. Ind. 2019, 105, 35–47. [CrossRef]
97. Patange, A.D.; Jegadeeshwaran, R. Application of bayesian family classifiers for cutting tool inserts health monitoring on CNC
milling. Int. J. Progn. Health Manag. 2020, 11. [CrossRef]
98. Pandita, P.; Ghosh, S.; Gupta, V.K.; Meshkov, A.; Wang, L. Application of Deep Transfer Learning and Uncertainty Quantification
for Process Identification in Powder Bed Fusion. ASME J. Risk Uncertain. Part B Mech. Eng. 2022, 8, 011106. [CrossRef]
99. Farahani, A.; Tohidi, H.; Shoja, A. An integrated optimization of quality control chart parameters and preventive maintenance
using Markov chain. Adv. Prod. Eng. Manag. 2019, 14, 5–14. [CrossRef]
100. El Haoud, N.; Bachiri, Z. Stochastic artificial intelligence benefits and supply chain management inventory prediction. In
Proceedings of the 2019 International Colloquium on Logistics and Supply Chain Management (LOGISTIQUA), Paris, France,
12–14 June 2019; pp. 1–5.
101. Feng, M.; Li, Y. Predictive Maintenance Decision Making Based on Reinforcement Learning in Multistage Production Systems.
IEEE Access 2022, 10, 18910–18921. [CrossRef]
102. Sobaszek, Ł.; Gola, A.; Kozłowski, E. Predictive scheduling with Markov chains and ARIMA models. Appl. Sci. 2020, 10, 6121.
[CrossRef]
103. Hofmann, T.; Schölkopf, B.; Smola, A.J. Kernel methods in machine learning. Ann. Stat. 2008, 36, 1171–1220. [CrossRef]
104. Gobert, C.; Reutzel, E.W.; Petrich, J.; Nassar, A.R.; Phoha, S. Application of supervised machine learning for defect detection
during metallic powder bed fusion additive manufacturing using high resolution imaging. Addit. Manuf. 2018, 21, 517–528.
[CrossRef]
105. McGregor, D.J.; Bimrose, M.V.; Shao, C.; Tawfick, S.; King, W.P. Using machine learning to predict dimensions and qualify diverse
part designs across multiple additive machines and materials. Addit. Manuf. 2022, 55, 102848. [CrossRef]
106. Kubik, C.; Knauer, S.M.; Groche, P. Smart sheet metal forming: Importance of data acquisition, preprocessing and transformation
on the performance of a multiclass support vector machine for predicting wear states during blanking. J. Intell. Manuf. 2022,
33, 259–282. [CrossRef]
107. Cortes, C.; Vapnik, V. Support-vector networks. Mach. Learn. 1995, 20, 273–297. [CrossRef]
108. Mika, S.; Ratsch, G.; Weston, J.; Scholkopf, B.; Mullers, K. Fisher discriminant analysis with kernels. In Proceedings of the
Neural Networks for Signal Processing IX: Proceedings of the 1999 IEEE Signal Processing Society Workshop (Cat. No.98TH8468),
Madison, WI, USA, 25 August 1999; pp. 41–48. [CrossRef]
109. Fukushima, K.; Miyake, S. Neocognitron: A self-organizing neural network model for a mechanism of visual pattern recognition.
In Competition and Cooperation in Neural Nets; Springer: Berlin/Heidelberg, Germany, 1982; pp. 267–285.
110. Rumelhart, D.E.; Hinton, G.E.; Williams, R.J. Learning representations by back-propagating errors. Nature 1986, 323, 533–536.
[CrossRef]
111. Hinton, G.E. Deep belief networks. Scholarpedia 2009, 4, 5947. [CrossRef]
112. Badmos, O.; Kopp, A.; Bernthaler, T.; Schneider, G. Image-based defect detection in lithium-ion battery electrode using
convolutional neural networks. J. Intell. Manuf. 2020, 31, 885–897. [CrossRef]
113. Ho, S.; Zhang, W.; Young, W.; Buchholz, M.; Al Jufout, S.; Dajani, K.; Bian, L.; Mozumdar, M. DLAM: Deep Learning Based Real-
Time Porosity Prediction for Additive Manufacturing Using Thermal Images of the Melt Pool. IEEE Access 2021, 9, 115100–115114.
[CrossRef]
114. Wen, L.; Li, X.; Gao, L.; Zhang, Y. A new convolutional neural network-based data-driven fault diagnosis method. IEEE Trans.
Ind. Electron. 2017, 65, 5990–5998. [CrossRef]
115. Al-Dulaimi, A.; Zabihi, S.; Asif, A.; Mohammadi, A. A multimodal and hybrid deep neural network model for remaining useful
life estimation. Comput. Ind. 2019, 108, 186–196. [CrossRef]
Appl. Sci. 2023, 13, 1903 29 of 32

116. Huang, J.; Segura, L.J.; Wang, T.; Zhao, G.; Sun, H.; Zhou, C. Unsupervised learning for the droplet evolution prediction and
process dynamics understanding in inkjet printing. Addit. Manuf. 2020, 35, 101197. [CrossRef]
117. Huang, J.; Chang, Q.; Arinez, J. Product completion time prediction using a hybrid approach combining deep learning and
system model. J. Manuf. Syst. 2020, 57, 311–322. [CrossRef]
118. Cohen, J.; Jiang, B.; Ni, J. Machine Learning for Diagnosis of Event Synchronization Faults in Discrete Manufacturing Systems. J.
Manuf. Sci. Eng. 2022, 144, 071006. [CrossRef]
119. Mujeeb, A.; Dai, W.; Erdt, M.; Sourin, A. One class based feature learning approach for defect detection using deep autoencoders.
Adv. Eng. Inform. 2019, 42, 100933. [CrossRef]
120. Kasim, N.; Nuawi, M.; Ghani, J.; Rizal, M.; Ngatiman, N.; Haron, C. Enhancing Clustering Algorithm with Initial Centroids in
Tool Wear Region Recognition. Int. J. Precis. Eng. Manuf. 2021, 22, 843–863. [CrossRef]
121. Djatna, T.; Alitu, I.M. An application of association rule mining in total productive maintenance strategy: An analysis and
modelling in wooden door manufacturing industry. Procedia Manuf. 2015, 4, 336–343. [CrossRef]
122. Chiang, L.H.; Colegrove, L.F. Industrial implementation of on-line multivariate quality control. Chemom. Intell. Lab. Syst. 2007,
88, 143–153. [CrossRef]
123. You, D.; Gao, X.; Katayama, S. WPD-PCA-based laser welding process monitoring and defects diagnosis by using FNN and SVM.
IEEE Trans. Ind. Electron. 2014, 62, 628–636. [CrossRef]
124. Moshat, S.; Datta, S.; Bandyopadhyay, A.; Pal, P. Optimization of CNC end milling process parameters using PCA-based Taguchi
method. Int. J. Eng. Sci. Technol. 2010, 2, 95–102. [CrossRef]
125. Mei, S.; Wang, Y.; Wen, G. Automatic fabric defect detection with a multi-scale convolutional denoising autoencoder network
model. Sensors 2018, 18, 1064. [CrossRef] [PubMed]
126. Maggipinto, M.; Beghi, A.; Susto, G.A. A Deep Convolutional Autoencoder-Based Approach for Anomaly Detection With
Industrial, Non-Images, 2-Dimensional Data: A Semiconductor Manufacturing Case Study. IEEE Trans. Autom. Sci. Eng. 2022.
[CrossRef]
127. Yang, Z.; Gjorgjevikj, D.; Long, J.; Zi, Y.; Zhang, S.; Li, C. Sparse autoencoder-based multi-head deep neural networks for
machinery fault diagnostics with detection of novelties. Chin. J. Mech. Eng. 2021, 34, 54. [CrossRef]
128. Cheng, R.C.; Chen, K.S. Ball bearing multiple failure diagnosis using feature-selected autoencoder model. Int. J. Adv. Manuf.
Technol. 2022, 120, 4803–4819. [CrossRef]
129. Ramamurthy, M.; Robinson, Y.H.; Vimal, S.; Suresh, A. Auto encoder based dimensionality reduction and classification using
convolutional neural networks for hyperspectral images. Microprocess. Microsyst. 2020, 79, 103280. [CrossRef]
130. Angelopoulos, A.; Michailidis, E.T.; Nomikos, N.; Trakadas, P.; Hatziefremidis, A.; Voliotis, S.; Zahariadis, T. Tackling faults in the
industry 4.0 era—A survey of machine-learning solutions and key aspects. Sensors 2019, 20, 109. [CrossRef]
131. de Lima, M.J.; Crovato, C.D.P.; Mejia, R.I.G.; da Rosa Righi, R.; de Oliveira Ramos, G.; da Costa, C.A.; Pesenti, G. HealthMon: An
approach for monitoring machines degradation using time-series decomposition, clustering, and metaheuristics. Comput. Ind.
Eng. 2021, 162, 107709. [CrossRef]
132. Song, W.; Wen, L.; Gao, L.; Li, X. Unsupervised fault diagnosis method based on iterative multi-manifold spectral clustering. IET
Collab. Intell. Manuf. 2019, 1, 48–55. [CrossRef]
133. Subramaniyan, M.; Skoogh, A.; Muhammad, A.S.; Bokrantz, J.; Johansson, B.; Roser, C. A generic hierarchical clustering approach
for detecting bottlenecks in manufacturing. J. Manuf. Syst. 2020, 55, 143–158. [CrossRef]
134. Srinivasan, M.; Moon, Y.B. A comprehensive clustering algorithm for strategic analysis of supply chain networks. Comput. Ind.
Eng. 1999, 36, 615–633. [CrossRef]
135. Das, J.N.; Tiwari, M.K.; Sinha, A.K.; Khanzode, V. Integrated warehouse assignment and carton configuration optimization using
deep clustering-based evolutionary algorithms. Expert Syst. Appl. 2023, 212, 118680. [CrossRef]
136. Stojanovic, L.; Dinic, M.; Stojanovic, N.; Stojadinovic, A. Big-data-driven anomaly detection in industry (4.0): An approach
and a case study. In Proceedings of the 2016 IEEE International Conference on Big Data (Big Data), Washington, DC, USA, 5–8
December 2016; pp. 1647–1652.
137. Saldivar, A.A.F.; Goh, C.; Li, Y.; Chen, Y.; Yu, H. Identifying smart design attributes for Industry 4.0 customization using a
clustering Genetic Algorithm. In Proceedings of the 2016 22nd International Conference on Automation and Computing (ICAC),
Colchester, UK, 7–8 September 2016; pp. 408–414.
138. Chen, W.C.; Tseng, S.S.; Wang, C.Y. A novel manufacturing defect detection method using association rule mining techniques.
Expert Syst. Appl. 2005, 29, 807–815. [CrossRef]
139. Shorten, C.; Khoshgoftaar, T.M. A survey on image data augmentation for deep learning. J. Big Data 2019, 6, 60. [CrossRef]
140. Iwana, B.K.; Uchida, S. An empirical survey of data augmentation for time series classification with neural networks. PLoS ONE
2021, 16, e0254841. [CrossRef]
141. Goodfellow, I.; Pouget-Abadie, J.; Mirza, M.; Xu, B.; Warde-Farley, D.; Ozair, S.; Courville, A.; Bengio, Y. Generative adversarial
networks. Commun. ACM 2020, 63, 139–144. [CrossRef]
142. Kingma, D.P.; Welling, M. Auto-encoding variational bayes. arXiv 2013, arXiv:1312.6114.
143. Wong, S.C.; Gatt, A.; Stamatescu, V.; McDonnell, M.D. Understanding data augmentation for classification: When to warp? In
Proceedings of the 2016 International Conference on Digital Image Computing: Techniques and Applications (DICTA), Gold
Coast, Australia, 30 November–2 December 2016; pp. 1–6.
Appl. Sci. 2023, 13, 1903 30 of 32

144. Berthelot, D.; Carlini, N.; Goodfellow, I.; Papernot, N.; Oliver, A.; Raffel, C.A. Mixmatch: A holistic approach to semi-supervised
learning. Adv. Neural Inf. Process. Syst. 2019, 32, 5049–5059.
145. Sohn, K.; Berthelot, D.; Carlini, N.; Zhang, Z.; Zhang, H.; Raffel, C.A.; Cubuk, E.D.; Kurakin, A.; Li, C.L. Fixmatch: Simplifying
semi-supervised learning with consistency and confidence. Adv. Neural Inf. Process. Syst. 2020, 33, 596–608.
146. Yang, X.; Song, Z.; King, I.; Xu, Z. A Survey on Deep Semi-supervised Learning. arXiv 2021, arXiv:2103.00550.
147. Sajjadi, M.; Javanmardi, M.; Tasdizen, T. Regularization with stochastic transformations and perturbations for deep semi-
supervised learning. Adv. Neural Inf. Process. Syst. 2016, 29, 1171–1179.
148. Tarvainen, A.; Valpola, H. Mean teachers are better role models: Weight-averaged consistency targets improve semi-supervised
deep learning results. Adv. Neural Inf. Process. Syst. 2017, 30, 1195–1204.
149. Li, X.; Jia, X.; Yang, Q.; Lee, J. Quality analysis in metal additive manufacturing with deep learning. J. Intell. Manuf. 2020,
31, 2003–2017. [CrossRef]
150. Zhao, B.; Zhang, X.; Zhan, Z.; Wu, Q.; Zhang, H. A Novel Semi-Supervised Graph-Guided Approach for Intelligent Health State
Diagnosis of a 3-PRR Planar Parallel Manipulator. IEEE/ASME Trans. Mechatron. 2022, 27, 4786–4797. [CrossRef]
151. Gilmer, J.; Schoenholz, S.S.; Riley, P.F.; Vinyals, O.; Dahl, G.E. Neural message passing for quantum chemistry. In Proceedings of
the International Conference on Machine Learning, Sydney, Australia, 6–11 August 2017; pp. 1263–1272.
152. Kipf, T.N.; Welling, M. Semi-supervised classification with graph convolutional networks. arXiv 2016, arXiv:1609.02907.
153. Serradilla, O.; Zugasti, E.; Ramirez de Okariz, J.; Rodriguez, J.; Zurutuza, U. Adaptable and explainable predictive maintenance:
Semi-supervised deep learning for anomaly detection and diagnosis in press machine data. Appl. Sci. 2021, 11, 7376. [CrossRef]
154. Song, J.; Lee, Y.C.; Lee, J. Deep generative model with time series-image encoding for manufacturing fault detection in die casting
process. J. Intell. Manuf. 2022, 1–14. [CrossRef]
155. Springenberg, J.T. Unsupervised and semi-supervised learning with categorical generative adversarial networks. arXiv 2015,
arXiv:1511.06390.
156. Salimans, T.; Goodfellow, I.; Zaremba, W.; Cheung, V.; Radford, A.; Chen, X. Improved techniques for training gans. Adv. Neural
Inf. Process. Syst. 2016, 29, 2234–2242.
157. Kingma, D.P.; Mohamed, S.; Jimenez Rezende, D.; Welling, M. Semi-supervised learning with deep generative models. Adv.
Neural Inf. Process. Syst. 2014, 27, 3581–3589.
158. Sutton, R.S.; Barto, A.G. Reinforcement Learning: An Introduction; MIT Press: Cambridge, MA, USA, 2018.
159. May, M.C.; Overbeck, L.; Wurster, M.; Kuhnle, A.; Lanza, G. Foresighted digital twin for situational agent selection in production
control. Procedia CIRP 2021, 99, 27–32. [CrossRef]
160. May, M.C.; Kiefer, L.; Kuhnle, A.; Stricker, N.; Lanza, G. Decentralized multi-agent production control through economic model
bidding for matrix production systems. Procedia Cirp 2021, 96, 3–8. [CrossRef]
161. Yao, M. Breakthrough Research In Reinforcement Learning From 2019. 2019. Available online: https://www.topbots.com/top-ai-
reinforcement-learning-research-papers-2019 (accessed on 1 September 2022).
162. Gao, R.X.; Wang, L.; Helu, M.; Teti, R. Big data analytics for smart factories of the future. CIRP Ann. 2020, 69, 668–692. [CrossRef]
163. Kozjek, D.; Vrabič, R.; Kralj, D.; Butala, P. Interpretative identification of the faulty conditions in a cyclic manufacturing process. J.
Manuf. Syst. 2017, 43, 214–224. [CrossRef]
164. Wen, Q.; Sun, L.; Yang, F.; Song, X.; Gao, J.; Wang, X.; Xu, H. Time series data augmentation for deep learning: A survey. arXiv
2020, arXiv:2002.12478.
165. Pan, S.J.; Yang, Q. A survey on transfer learning. IEEE Trans. Knowl. Data Eng. 2009, 22, 1345–1359. [CrossRef]
166. Zhang, H.; Cisse, M.; Dauphin, Y.N.; Lopez-Paz, D. mixup: Beyond empirical risk minimization. arXiv 2017, arXiv:1710.09412.
167. Bao, J.; Chen, D.; Wen, F.; Li, H.; Hua, G. CVAE-GAN: Fine-grained image generation through asymmetric training. In
Proceedings of the IEEE International Conference on Computer Vision, Venice, Italy, 22–29 October 2017; pp. 2745–2754.
168. Yoon, J.; Jarrett, D.; Van der Schaar, M. Time-series generative adversarial networks. Adv. Neural Inf. Process. Syst. 2019, 32,
5508–5518.
169. McMahan, B.; Moore, E.; Ramage, D.; Hampson, S.; y Arcas, B.A. Communication-efficient learning of deep networks from
decentralized data. In Proceedings of the Artificial Intelligence and Statistics, Fort Lauderdale, FL, USA, 20–22 April 2017;
pp. 1273–1282.
170. Cheng, Y.; Wang, D.; Zhou, P.; Zhang, T. Model compression and acceleration for deep neural networks: The principles, progress,
and challenges. IEEE Signal Process. Mag. 2018, 35, 126–136. [CrossRef]
171. Gou, J.; Yu, B.; Maybank, S.J.; Tao, D. Knowledge distillation: A survey. Int. J. Comput. Vis. 2021, 129, 1789–1819. [CrossRef]
172. Hinton, G.; Vinyals, O.; Dean, J. Distilling the knowledge in a neural network. arXiv 2015, arXiv:1503.02531.
173. Schlimmer, J.C.; Granger, R.H. Incremental learning from noisy data. Mach. Learn. 1986, 1, 317–354. [CrossRef]
174. Gama, J.; Žliobaitė, I.; Bifet, A.; Pechenizkiy, M.; Bouchachia, A. A survey on concept drift adaptation. ACM Comput. Surv. 2014,
46, 44. [CrossRef]
175. Baier, L.; Jöhren, F.; Seebacher, S. Challenges in the Deployment and Operation of Machine Learning in Practice. In Proceedings
of the ECIS 2019 27th European Conference on Information Systems, Stockholm, Sweden, 8–14 June 2019.
176. Canbek, G. Gaining insights in datasets in the shade of “garbage in, garbage out” rationale: Feature space distribution fitting.
Wiley Interdiscip. Rev. Data Min. Knowl. Discov. 2022, 12, e1456. [CrossRef]
Appl. Sci. 2023, 13, 1903 31 of 32

177. Moges, T.; Yang, Z.; Jones, K.; Feng, S.; Witherell, P.; Lu, Y. Hybrid modeling approach for melt-pool prediction in laser powder
bed fusion additive manufacturing. J. Comput. Inf. Sci. Eng. 2021, 21, 050902. [CrossRef]
178. Colledani, M., Statistical Process Control. In CIRP Encyclopedia of Production Engineering; Laperrière, L., Reinhart, G., Eds.;
Springer: Berlin/Heidelberg, Germany, 2014; pp. 1150–1157. [CrossRef]
179. Abdar, M.; Pourpanah, F.; Hussain, S.; Rezazadegan, D.; Liu, L.; Ghavamzadeh, M.; Fieguth, P.; Cao, X.; Khosravi, A.;
Acharya, U.R.; et al. A review of uncertainty quantification in deep learning: Techniques, applications and challenges. Inf. Fusion
2021, 76, 243–297. [CrossRef]
180. Yong, B.X.; Brintrup, A. Multi Agent System for Machine Learning Under Uncertainty in Cyber Physical Manufacturing System.
In Service Oriented, Holonic and Multi-Agent Manufacturing Systems for Industry of the Future; Borangiu, T., Trentesaux, D., Leitão, P.,
Giret Boggino, A., Botti, V., Eds.; Studies in Computational Intelligence; Springer International Publishing: Cham, Switzerland,
2020; Volume 853, pp. 244–257. [CrossRef]
181. Tavazza, F.; DeCost, B.; Choudhary, K. Uncertainty Prediction for Machine Learning Models of Material Properties. ACS Omega
2021, 6, 32431–32440. [CrossRef]
182. Arkov, V. Uncertainty Estimation in Machine Learning. arXiv 2022. [CrossRef]
183. Zhang, B. Data-Driven Uncertainty Analysis in Neural Networks with Applications to Manufacturing Process Monitoring. Ph.D.
Thesis, Purdue University Graduate School, West Lafayette, IN, USA, 2021. [CrossRef]
184. Zhang, B.; Shin, Y.C. A probabilistic neural network for uncertainty prediction with applications to manufacturing process
monitoring. Appl. Soft Comput. 2022, 124, 108995. [CrossRef]
185. Lee, S.; Kim, S.B. Time-adaptive support vector data description for nonstationary process monitoring. Eng. Appl. Artif. Intell.
2018, 68, 18–31. [CrossRef]
186. Gaikwad, A.; Yavari, R.; Montazeri, M.; Cole, K.; Bian, L.; Rao, P. Toward the digital twin of additive manufacturing: Integrating
thermal simulations, sensing, and analytics to detect process faults. IISE Trans. 2020, 52, 1204–1217. [CrossRef]
187. Zhang, C.J.; Zhang, Y.C.; Han, Y. Industrial cyber-physical system driven intelligent prediction model for converter end carbon
content in steelmaking plants. J. Ind. Inf. Integr. 2022, 28, 100356. [CrossRef]
188. Ning, F.; Shi, Y.; Cai, M.; Xu, W.; Zhang, X. Manufacturing cost estimation based on the machining process and deep-learning
method. J. Manuf. Syst. 2020, 56, 11–22. [CrossRef]
189. Westphal, E.; Seitz, H. Machine learning for the intelligent analysis of 3D printing conditions using environmental sensor data to
support quality assurance. Addit. Manuf. 2022, 50, 102535. [CrossRef]
190. Qin, J.; Wang, Y.; Ding, J.; Williams, S. Optimal droplet transfer mode maintenance for wire+ arc additive manufacturing (WAAM)
based on deep learning. J. Intell. Manuf. 2022, 33, 2179–2191. [CrossRef]
191. Lapointe, S.; Guss, G.; Reese, Z.; Strantza, M.; Matthews, M.; Druzgalski, C. Photodiode-based machine learning for optimization
of laser powder bed fusion parameters in complex geometries. Addit. Manuf. 2022, 53, 102687. [CrossRef]
192. Zhang, T.; Zhang, C.; Hu, T. A robotic grasp detection method based on auto-annotated dataset in disordered manufacturing
scenarios. Robot. Comput. Integr. Manuf. 2022, 76, 102329. [CrossRef]
193. Singh, S.A.; Desai, K. Automated surface defect detection framework using machine vision and convolutional neural networks. J.
Intell. Manuf. 2022, 1–17. [CrossRef]
194. Duan, J.; Hu, C.; Zhan, X.; Zhou, H.; Liao, G.; Shi, T. MS-SSPCANet: A powerful deep learning framework for tool wear
prediction. Robot. Comput. Integr. Manuf. 2022, 78, 102391. [CrossRef]
195. Gao, K.; Chen, H.; Zhang, X.; Ren, X.; Chen, J.; Chen, X. A novel material removal prediction method based on acoustic sensing
and ensemble XGBoost learning algorithm for robotic belt grinding of Inconel 718. Int. J. Adv. Manuf. Technol. 2019, 105, 217–232.
[CrossRef]
196. Gawade, V.; Singh, V.; Guo, W. Leveraging simulated and empirical data-driven insight to supervised-learning for porosity
prediction in laser metal deposition. J. Manuf. Syst. 2022, 62, 875–885. [CrossRef]
197. Aminzadeh, M.; Kurfess, T.R. Online quality inspection using Bayesian classification in powder-bed additive manufacturing
from high-resolution visual camera images. J. Intell. Manuf. 2019, 30, 2505–2523. [CrossRef]
198. Priore, P.; Ponte, B.; Puente, J.; Gómez, A. Learning-based scheduling of flexible manufacturing systems using ensemble methods.
Comput. Ind. Eng. 2018, 126, 282–291. [CrossRef]
199. Guo, S.; Chen, M.; Abolhassani, A.; Kalamdani, R.; Guo, W.G. Identifying manufacturing operational conditions by physics-based
feature extraction and ensemble clustering. J. Manuf. Syst. 2021, 60, 162–175. [CrossRef]
200. Kim, J.; Ko, J.; Choi, H.; Kim, H. Printed circuit board defect detection using deep learning via a skip-connected convolutional
autoencoder. Sensors 2021, 21, 4968. [CrossRef]
201. Jakubowski, J.; Stanisz, P.; Bobek, S.; Nalepa, G.J. Anomaly Detection in Asset Degradation Process Using Variational Autoencoder
and Explanations. Sensors 2021, 22, 291. [CrossRef]
202. Sarita, K.; Devarapalli, R.; Kumar, S.; Malik, H.; Garcia Marquez, F.P.; Rai, P. Principal component analysis technique for early
fault detection. J. Intell. Fuzzy Syst. 2022, 42, 861–872. [CrossRef]
203. Zheng, X.; Wang, H.; Chen, J.; Kong, Y.; Zheng, S. A generic semi-supervised deep learning-based approach for automated
surface inspection. IEEE Access 2020, 8, 114088–114099. [CrossRef]
204. Zhang, W.; Lang, J. Semi-supervised training for positioning of welding seams. Sensors 2021, 21, 7309. [CrossRef]
Appl. Sci. 2023, 13, 1903 32 of 32

205. Chen, C.; Liu, Y.; Kumar, M.; Qin, J.; Ren, Y. Energy consumption modelling using deep learning embedded semi-supervised
learning. Comput. Ind. Eng. 2019, 135, 757–765. [CrossRef]
206. Jun, J.h.; Chang, T.W.; Jun, S. Quality prediction and yield improvement in process manufacturing based on data analytics.
Processes 2020, 8, 1068. [CrossRef]
207. Shim, J.; Cho, S.; Kum, E.; Jeong, S. Adaptive fault detection framework for recipe transition in semiconductor manufacturing.
Comput. Ind. Eng. 2021, 161, 107632. [CrossRef]
208. Qiu, C.; Li, K.; Li, B.; Mao, X.; He, S.; Hao, C.; Yin, L. Semi-supervised graph convolutional network to predict position-and
speed-dependent tool tip dynamics with limited labeled data. Mech. Syst. Signal Process. 2022, 164, 108225. [CrossRef]
209. Guo, Y.; Lu, W.F.; Fuh, J.Y.H. Semi-supervised deep learning based framework for assessing manufacturability of cellular
structures in direct metal laser sintering process. J. Intell. Manuf. 2021, 32, 347–359. [CrossRef]
210. Okaro, I.A.; Jayasinghe, S.; Sutcliffe, C.; Black, K.; Paoletti, P.; Green, P.L. Automatic fault detection for laser powder-bed fusion
using semi-supervised machine learning. Addit. Manuf. 2019, 27, 42–53. [CrossRef]
211. Lee, H.; Kim, H. Semi-supervised multi-label learning for classification of wafer bin maps with mixed-type defect patterns. IEEE
Trans. Semicond. Manuf. 2020, 33, 653–662. [CrossRef]
212. Liu, J.; Song, K.; Feng, M.; Yan, Y.; Tu, Z.; Zhu, L. Semi-supervised anomaly detection with dual prototypes autoencoder for
industrial surface inspection. Opt. Lasers Eng. 2021, 136, 106324. [CrossRef]
213. Verstraete, D.; Droguett, E.; Modarres, M. A deep adversarial approach based on multi-sensor fusion for semi-supervised
remaining useful life prognostics. Sensors 2019, 20, 176. [CrossRef]
214. Souza, M.L.H.; da Costa, C.A.; de Oliveira Ramos, G.; da Rosa Righi, R. A feature identification method to explain anomalies in
condition monitoring. Comput. Ind. 2021, 133, 103528. [CrossRef]
215. Lee, Y.H.; Lee, S. Deep reinforcement learning based scheduling within production plan in semiconductor fabrication. Expert
Syst. Appl. 2022, 191, 116222. [CrossRef]
216. Marchesano, M.G.; Guizzi, G.; Santillo, L.C.; Vespoli, S. A deep reinforcement learning approach for the throughput control of a
flow-shop production system. IFAC-PapersOnLine 2021, 54, 61–66. [CrossRef]
217. Yang, H.; Li, W.; Wang, B. Joint optimization of preventive maintenance and production scheduling for multi-state production
systems based on reinforcement learning. Reliab. Eng. Syst. Saf. 2021, 214, 107713. [CrossRef]
218. Schneckenreither, M.; Haeussler, S.; Peiró, J. Average reward adjusted deep reinforcement learning for order release planning in
manufacturing. Knowl.-Based Syst. 2022, 247, 108765. [CrossRef]
219. Tsai, Y.T.; Lee, C.H.; Liu, T.Y.; Chang, T.J.; Wang, C.S.; Pawar, S.J.; Huang, P.H.; Huang, J.H. Utilization of a reinforcement learning
algorithm for the accurate alignment of a robotic arm in a complete soft fabric shoe tongues automation process. J. Manuf. Syst.
2020, 56, 501–513. [CrossRef]
220. Klar, M.; Glatt, M.; Aurich, J.C. An implementation of a reinforcement learning based algorithm for factory layout planning.
Manuf. Lett. 2021, 30, 1–4. [CrossRef]
221. Huang, J.; Chang, Q.; Arinez, J. Deep reinforcement learning based preventive maintenance policy for serial production lines.
Expert Syst. Appl. 2020, 160, 113701. [CrossRef]
222. Zhang, H.; Peng, Q.; Zhang, J.; Gu, P. Planning for automatic product assembly using reinforcement learning. Comput. Ind. 2021,
130, 103471. [CrossRef]
223. Kuhnle, A.; May, M.C.; Schaefer, L.; Lanza, G. Explainable reinforcement learning in production control of job shop manufacturing
system. Int. J. Prod. Res. 2021, 60, 5812–5834. [CrossRef]
224. Valet, A.; Altenmüller, T.; Waschneck, B.; May, M.C.; Kuhnle, A.; Lanza, G. Opportunistic maintenance scheduling with deep
reinforcement learning. J. Manuf. Syst. 2022, 64, 518–534. [CrossRef]
225. Huang, J.; Su, J.; Chang, Q. Graph neural network and multi-agent reinforcement learning for machine-process-system integrated
control to optimize production yield. J. Manuf. Syst. 2022, 64, 81–93. [CrossRef]
226. Zimmerling, C.; Poppe, C.; Stein, O.; Kärger, L. Optimisation of manufacturing process parameters for variable component
geometries using reinforcement learning. Mater. Des. 2022, 214, 110423. [CrossRef]
227. Guo, F.; Zhou, X.; Liu, J.; Zhang, Y.; Li, D.; Zhou, H. A reinforcement learning decision model for online process parameters
optimization from offline data in injection molding. Appl. Soft Comput. J. 2019, 85, 105828. [CrossRef]
228. Hofmann, C.; Liu, X.; May, M.; Lanza, G. Hybrid Monte Carlo tree search based multi-objective scheduling. Prod. Eng. 2022, 17,
133–144. [CrossRef]

Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.

You might also like