Next Article in Journal
An Indoor Fire Detection Method Based on Multi-Sensor Fusion and a Lightweight Convolutional Neural Network
Next Article in Special Issue
Residual Shrinkage ViT with Discriminative Rebalancing Strategy for Small and Imbalanced Fault Diagnosis
Previous Article in Journal
Deep Reinforcement Learning-Based Power Allocation for Minimizing Age of Information and Energy Consumption in Multi-Input Multi-Output and Non-Orthogonal Multiple Access Internet of Things Systems
Previous Article in Special Issue
Modelling and Analysis of Expansion Joints’ Effect on Dynamic Performance of Railway Rigid Overhead System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Fault Detection Approach Based on One-Sided Domain Adaptation and Generative Adversarial Networks for Railway Door Systems

1
Integrated Vehicle Health Management Centre, Cranfield University, Cranfield MK43 0AL, UK
2
Centre for Life-Cycle Engineering and Management, Cranfield University, Cranfield MK43 0AL, UK
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(24), 9688; https://doi.org/10.3390/s23249688
Submission received: 3 November 2023 / Revised: 30 November 2023 / Accepted: 5 December 2023 / Published: 7 December 2023

Abstract

:
Fault detection using the domain adaptation technique is one of the more promising methods of solving the domain shift problem, and has therefore been intensively investigated in recent years. However, the domain adaptation method still has elements of impracticality: firstly, domain-specific decision boundaries are not taken into consideration, which often results in poor performance near the class boundary; and secondly, information on the source domain needs to be exploited with priority over information on the target domain, as the source domain can provide a rich dataset. Thus, the real-world implementations of this approach are still scarce. In order to address these issues, a novel fault detection approach based on one-sided domain adaptation for real-world railway door systems is proposed. An anomaly detector created using label-rich source domain data is used to generate distinctive source latent features, and the target domain features are then aligned toward the source latent features in a one-sided way. The performance and sensitivity analyses show that the proposed method is more accurate than alternative methods, with an F1 score of 97.9%, and is the most robust against variation in the input features. The proposed method also bridges the gap between theoretical domain adaptation research and tangible industrial applications. Furthermore, the proposed approach can be applied to conventional railway components and various electro-mechanical actuators. This is because the motor current signals used in this study are primarily obtained from the controller or motor drive, which eliminates the need for extra sensors.

1. Introduction

Fault detection plays a vital role in maintenance tasks within the railway sector, and has been defined as “the detection of a fault within a prescribed time by a safety mechanism” [1]. Although railway machinery represents a complex industrial system with a huge variety of components, a train door is a vital subsystem that can lead to service interruptions or failures, resulting in higher operational and maintenance expenses. One report has indicated that the door system accounts for 30–60% of all malfunctions in railway vehicles [2]. To avoid such failures, predictive maintenance using data-driven methods has recently gained interest from researchers, owing to the vast quantities of monitoring data accessible.
Data-driven approaches for fault detection include traditional machine learning (ML) and deep learning (DL) approaches. Traditional ML approaches necessitate multiple steps such as data preprocessing and feature extraction prior to model development; however, manual feature extraction requires specialised domain expertise, which complicates the use of traditional machine learning methods. In contrast, DL methods allow for the development of fault detection models without the need for manually crafted features by employing a deep network architecture. This represents a notable advantage over traditional ML techniques.
Fault detection methods based on deep learning can be divided into supervised and unsupervised learning approaches. Supervised DL methods require datasets with labels to enable the training of the model. A significant proportion of prior research in the area of fault detection has focused on supervised DL methods, including deep neural networks (DNNs) [3], two-dimensional convolutional neural networks (2D CNNs) [4], one-dimensional convolutional neural networks (1D CNNs) [5], gated recurrent units (GRUs) [6], and long short-term memory (LSTM) [7]. However, the need for ample labelled datasets represents a major limitation of supervised methods: faulty data are often scarce, due to the use of conservative maintenance schedules to prevent major incidents. In addition, the imbalance between faulty and healthy data is also problematic when building a classifier. Unlike supervised approaches, unsupervised DL methods do not need datasets with labels, and the aim is to extract the relevant characteristics of the input data. Previous research based on unsupervised learning approaches has included stacked autoencoders [8], denoising autoencoders [9], sparse autoencoders [10], variational autoencoders [11], and deep belief networks [12].
Despite the successful outcomes of previous research, one significant drawback of a data-driven fault detection approach is that the fault detection performance may be considerably degraded when the model is applied to actual acquired data rather than training data. The assumption underlying traditional ML and DL approaches is that the distribution of the test data is identical to that of the training data; however, these distributions may differ due to the different operating conditions, components, and detailed specifications of the machinery. This is known as the domain shift problem [13], which refers to the discrepancy in the feature distribution between two domains. In this case, the training and test data are assumed to be the source and target domains, respectively. Due to the discrepancy between the source and target domains, the accuracy of fault detection in actual industrial data may be worse than anticipated. If the plan is to acquire many types of data beforehand, under different operating conditions and for different components, then the training of the model may be expensive and demanding, meaning that a huge dataset would be required. This strategy is therefore impractical, given that the availability of datasets that include sufficient faulty samples is always limited in industry.
The domain adaptation (DA) technique offers a promising solution for addressing the domain shift problem. DA is an area of machine learning where models are designed to execute tasks in a target domain, using knowledge gained from a similar but distinct source domain. The aim is to address the challenge posed by the distribution shift between the source and target domains. Research into DA has evolved over the years in several areas of study, including computer vision, healthcare, speech recognition, and fault detection. When applied in the context of fault detection, the aim of DA is to align the source and target domain distributions. The aligned source and target features are then used to build a fault detection model. As a result, faults can be detected with almost the same accuracy for both the source and target domains.
Fault detection based on DA can be categorised at the methodological level into network-based, instance-based, mapping-based, and adversarial-based approaches [14]. Network-based DA involves the direct transfer of certain network parameters that have been pre-trained in the source domain to another model in the target domain, as partial network parameters. The fine-tuning of the network parameters is then conducted using a limited set of labelled data from the target domain [15,16]. However, sufficient labelled target domain datasets are necessary for this method, which are often unavailable in the context of fault detection. Instance-based DA involves adjusting the weights of instances in the source domain to assist the classifier in label prediction or using instance statistics to bring the target domain into alignment. These methods include DNNs with a batch normalisation layer (BN) [17] and adaptive batch normalisation [18,19]. However, in order to train the BN layer and set appropriate parameters, a certain amount of normal and faulty target samples is required beforehand.
The aim of mapping-based DA is to project the original features from both the source and target domains into a new feature space, where the two domain features are aligned using a feature extractor. There are many examples of fault detection using mapping-based DA, including Kullback–Leibler divergence [20], correlation alignment (CORAL) [21], maximum mean discrepancy (MMD) [22,23], multi-kernel MMD [24,25], and joint distribution adaptation [26]. In contrast, adversarial-based DA uses an adversarial method in which a domain discriminator is used to minimise the discrepancy in the feature distribution between the source and target domains created by a feature extractor. The adversarial network architecture is called a generative adversarial network (GAN). This consists of a pair of networks that are combined to form a generative system, and was initially proposed by Ian Goodfellow in 2014 [27]. When carrying out domain adaptation with a GAN, the generator is typically employed to map the raw features from both the source and target domains to a new latent feature space, where the alignment of the feature distributions can be achieved. A fault detection model can be built using the aligned features in the latent feature space for both the source and target domains. Examples of adversarial-based DA include the domain adversarial neural network (DANN) [28,29], the domain adversarial transfer network [30], and the Wasserstein distance-based deep transfer network [31].
Adversarial-based DA can be integrated with mapping-based DA by employing both the adversarial discriminator and an appropriate objective function to minimise the discrepancy between the two domains. Research using both adversarial and mapping-based DA is being widely conducted in relation to fault detection, due to its strong ability to align two domains where only unlabelled target samples are available [31,32]. However, these methods still have some impractical aspects, which can be research gaps, as follows:
  • In a GAN, the feature generator does not take domain-specific decision boundaries into consideration, as the model simply tries to fool the discriminator [33]. This results in poor performance in terms of detecting faulty samples near the class boundary.
  • In general, the DA method causes both the source and target domain to be the same in the latent space, meaning that they are treated equally. However, the source domain information needs to be considered with a higher priority than the target domain, as the source domain tends to be a rich dataset that includes more faulty samples than the target domain under actual industrial conditions. The reason for this is that the fault detection model is initially built with a focus on the specific machinery, followed by thorough validation in order to make the model applicable to the actual industrial setting. The model is then applied to another domain.
  • Ensuring that DA techniques are robust and reliable for real-world applications is challenging if the method is completely unsupervised. This is because the degree of similarity between the two domains that is required in order to be able to apply the DA method successfully is unclear. However, reliability is crucial for fault detection to avoid catastrophic incidents; hence, real-world implementations of the DA technique are still scarce.
These are major challenges, and few studies can be found that have attempted to overcome these hurdles. In order to tackle these issues, a novel fault detection approach for railway door systems is proposed, based on one-sided DA using GANs. In this study, the source and target domains consist of data from a linear actuator test rig and a real-world railway door, respectively. First, an anomaly detector and a feature generator are trained using label-rich source domain data to generate distinctive source latent features. Next, the target domain data are aligned with the source latent features in a one-sided way. The proposed method enables the faulty target samples to be aligned with the source samples and to be detected accurately by the same anomaly detector, which is built based on rich source data. To the best of our knowledge, this paper is the first to introduce a fault detection method utilising DA specifically for railway door systems. The main contributions of the paper can be summarised as follows:
  • A fault detection approach is proposed based on one-sided DA with GANs, which can be used for real-world railway door systems.
  • The proposed one-sided DA from the target to the source domain enables the normal and faulty samples in the target domain to be detected using the same fault detection model, which is trained on a rich source dataset.
  • Our approach ensures that the two domains can be aligned, despite the low level of similarity between different components, using only a few faulty target samples.
  • The proposed method is not only the most accurate and robust among comparative models but also bridges the gap between theoretical domain adaptation research and tangible industrial applications.
  • The proposed approach can also be applied to conventional railway components and various electro-mechanical actuators. This is because the motor current signals considered in this study are primarily obtained from the controller or motor drive, thus eliminating the need for extra sensors.
The remainder of this article is organised as follows. Section 2 introduces the dataset and the proposed methodology. Some results and a discussion are given in Section 3. Finally, Section 4 concludes this article.

2. Materials and Methods

2.1. Dataset

2.1.1. Linear Actuator Experimental Dataset

The primary component of the test rig was a ball screw mechanism, featuring a threaded shaft with a helical raceway that facilitated the movement of the bearing balls contained within the nut [34,35]. Different loads were produced by connecting a secondary actuator. The actuators were linked via a load cell, which supplied feedback to the controller. This setup allowed for the generation of various operating conditions by altering the load setpoint. In this case, the load setpoints used were 196.13 N, 392.3 N, and −392.3 N. Three different types of fault were introduced, with increasing severity: a lack of lubrication, spalling, and backlash. The tests were carried out using two types of motion profiles: trapezoidal (for constant speed) and sinusoidal (for smooth acceleration and deceleration). A 3D representation of the test rig, along with a side view, is presented in Figure 1. More comprehensive information about the test rig and the introduced faults is available in [34], and the raw data can be accessed and downloaded from [36].
In our research, trapezoidal motion profiles were selected to construct the model, as railway door systems typically exhibit relatively constant speed profiles, as explained in Section 2.1.2. The measurement of the position and current signals involved both extension and retraction processes, as illustrated in Figure 2. The current signals specific to the extension operation were extracted and used to build the model, as this operation represented the closing mechanism of the railway door systems considered in this research, as explained in detail in Section 2.1.2. To reduce noise, a low-pass filter with a window of 0.15 s was applied. The current profiles that indicated a lack of lubrication were selected as the faulty current signals. Both normal and faulty current profiles are shown in Figure 3, where the various characteristics of the faulty signals can be observed; for instance, although there is some overshoot in the normal profiles, this overshoot is reduced in the faulty profiles. The dataset comprises three distinct loading conditions, which are all considered under the same class label. For example, normal profiles from these three loading conditions, as illustrated in Figure 3, are classified as the normal class, and the same categorisation applies conversely.

2.1.2. Operational Datasets for Railway Door Systems

This research used extensive real-world datasets collected from railway door systems. The focus was on an electric door system consisting of a voltage power source, a DC motor, a door control unit (DCU), a transmission system, and the door leaves. The DC motor, energised by the voltage source and regulated by the DCU, delivered the required shaft angular velocity and torque, which were then conveyed to the transmission system to enable the door leaves to move in a predetermined fashion [37]. The current signal from the door was gathered via the communication port of the DCU at a frequency of 50 Hz. A low-pass filter with a 0.25 s window, equivalent to five consecutive measurement periods, was implemented to minimise the noise in the current signals.
Figure 4 presents some examples of signal profiles for both the opening and closing operations. In the opening profile, there is a steady increase in both the speed and current up to a peak, and then a gentle curve and a decline to zero. The closing profile exhibits a pattern similar to the opening profile, but with two notable differences in the current: firstly, the peak current for the closing process is lower than that for the opening process, and secondly, there is a sharp change towards the end of the closing profile. This is accompanied by a minor increase in the speed, which enables the door to reach its fully closed position, where the locking mechanism can be activated [38]. It should be noted that specific types of faults cannot be identified in this dataset [37]. The experimental current signals from the linear actuator in the three fault modes (lack of lubrication, spalling, and backlash) were compared with the faulty signals from railway door systems. However, none of these fault modes showed a similarity to the faulty signals observed in railway door operations. This suggests that the faulty behaviour detected in train doors could be attributed to the multiple fault modes of the numerous components in the train door system.
In this study, current signals from closing operations were employed to detect faults. Examples of normal and faulty signals for the closing operation are shown in Figure 4. In the normal current signal, there are flat curves between 3.2 s and 4.0 s, in contrast to the negative peaks and variations observed in the plot of the faulty data. It is noteworthy that the characteristics of the faults in these door systems are different from those observed in a linear actuator test rig, as explained in Section 2.1.1. Although a specific faulty mode is used as an example, it is noteworthy that the proposed method aims to be universally applicable across various types of fault modes, as it does not rely on assumptions specific to any particular fault mode.

2.2. Proposed Methodology

The workflow for the proposed methodology for railway door systems, based on one-sided DA with a GAN, is shown in Figure 5. The workflow is divided into two procedures, marked Steps 1 and 2. The source and target domain datasets can be described as follows:
D s = x i s ,   y i s i = 1 n s
D t = x i t ,   y i t i = 1 n t
where Ds and Dt are the source and target domain datasets, xs and xt are the source and target feature vectors, ys and yt are the source and target labels (which are used to categorise the data into two states: normal or faulty), and ns and nt are the numbers of the source and target samples, respectively.

2.2.1. Step 1: Train a Feature Extractor and an Anomaly Detector on the Source Dataset

A feature extractor and an anomaly detector, with the network architectures and hyperparameters given in Table 1 and Table 2, were trained using only the source domain dataset. In order to train the models, a binary cross-entropy (BCE) loss was used as a loss function for both the feature extractor and the anomaly detector. The BCE loss and the optimisation objective are expressed as follows:
L a = 1 n s i = 1 n s y i s l n f a f e x i s + 1 y i s l n 1 f a f e x i s
θ e * , θ a * = a r g m i n θ e * , θ a * L a
where La is a loss function for both the feature extractor and anomaly detector; fe and fa are functions of the feature extractor and the anomaly detector, respectively, which are parameterised by θe and θa; and θ* is the optimised value of θ. Notably, fe is a mapping function, whereas fa is a classifier. As shown in Equation (4), the optimisation objective for both the feature extractor and the anomaly detector is to minimise La. In view of this optimisation objective, the feature extractor is forced to generate features that are separable by the anomaly detector. It is therefore assumed that faulty samples can be detected by the anomaly detector with high accuracy, meaning that the extracted features of the normal and faulty source samples should be distinct in the latent space. Hence, these features are useful in classifying samples into normal or faulty classes.

2.2.2. Step 2: One-Sided DA from the Target Domain to the Source Domain

Once Step 1 is complete, a feature generator and a discriminator are trained in an adversarial manner. The loss function for the feature generator includes three loss items: MMD loss, a fault detection loss, and an adversarial loss, as follows:
L g = L M M D + L F D + L A D
θ g * = a r g m i n θ g L g
where LMMD, LFD, and LAD are the MMD loss, the fault detection loss, and the adversarial loss, respectively. θg is a learnable parameter of the feature generator fg.
The MMD is a statistical measure that is used to quantify the dissimilarity between two distributions, and was initially introduced by Gretton et al. [39]. The concept of MMD is closely related to kernel methods, and it has been used in various areas of machine learning, including domain adaptation and generative modelling. The MMD is defined as a squared distance in the reproducing kernel Hilbert space (RKHS), and can be expressed as follows:
L M M D = M M D f e x s ,   f g x t = E f e x s ~ P φ f e x s E f g x t ~ Q φ f g x t H 2
k x , y = φ x , φ y H = e x p x y 2 σ
E f e x s ~ P φ f e x s E f g x t ~ Q φ f g x t H 2 = E f e x s , f e x s ~ P k f e x s , f e x s + E f g x t , f g x t ~ Q k f g x t , f g x t 2 E f e x s ~ P , f g x t ~ Q , k f e x s , f g x t
where fe(xs) and fg(xt) are the latent features from the distributions P and Q, H represents RKHS using kernel k, and φ is a mapping function to RKHS. A radial basis function (RBF) kernel, also known as a Gaussian kernel, is chosen as k in this research, and σ is set to one. The RBF kernel enables computing the kernel function k directly without explicitly knowing the form of φ, which is known as the kernel trick.
It is noteworthy that the MMD loss measures the discrepancy between the two latent feature distributions of the source and target domain, which are fe(xs) and fg(xt). The latent features of the source domain are extracted by the feature extractor, which is built in Step 1 and fixed during the training process in Step 2. This means that the source latent features remain invariant in Step 2. In contrast, the feature generator is trained to generate target latent features that are as identical as possible to the source latent features, in order to minimise the MMD loss.
The fault detection loss in Step 2, on the other hand, is calculated using the BCE loss in Equation (10), using the anomaly detector with only target data, as shown in the following equation:
L F D = 1 n t i = 1 n t y i t l n f a f g x i t + 1 y i t l n 1 f a f g x i t
The anomaly detector fa built in Step 1 is employed and fixed while training the feature generator, meaning that only the feature generator is trained to minimise LFD. Hence, once training has been conducted, the anomaly detector should also accurately classify normal and faulty target samples.
The discriminator model is a classifier, and aims to distinguish whether the input sample is a source or target sample. It is trained using both the latent source and target features. A unified dataset D′ is defined, which encompasses both the source and target latent features, as follows:
D = z j ,   l j j = 1 n s + n t
z j ,   l j = f e x j s ,   l s   i f   j     n s ,   w h e r e   l s   d e n o t e s   t h e   l a b e l   s o u r c e f g x j t ,   l t   i f   j > n s ,   w h e r e   l t   d e n o t e s   t h e   l a b e l   t a r g e t
The discriminator loss for the discriminator and the adversarial loss for the generator are expressed as in the following equations:
L D = 1 n s + n t j = 1 n s + n t l j l n f d z j + 1 l j l n 1 f d z j
L A D = 1 n s + n t j = 1 n s + n t l j l n 1 f d z j + 1 l j l n f d z j
θ d * = a r g m i n θ d L D
where LD is the discriminator loss, and fd is a discriminator function parameterised by θd. The objective of the discriminator is to minimise the BCE loss for the discriminator, as described in Equation (15). In contrast, Equations (6) and (14) show that the feature generator is trained to maximise the discriminator loss, as LD should be a maximum when LAD is minimised. Thus, the optimisation goal for the feature generator is to fool the discriminator, whereas the discriminator is trained to distinguish between the source and target samples in an adversarial manner. The purpose of employing the adversarial loss is to ensure that the latent source and target features are identical to each other.
To summarise, the purposes of each loss item described in Equation (5) are as follows:
-
LMMD forces the feature generator to create target latent features that are as identical as possible to the source latent features as a DA capability.
-
LFD enables normal and faulty target samples to be distinctive and to be detected by the same anomaly detector trained on a rich source domain dataset.
-
LAD ensures that the latent source and target features are identical.
The model built using the proposed method was a simple neural network-based model, as shown in Table 1 and Table 2; however, any other DL model architecture, such as CNN and LSTM, could be employed as long as they have equivalent loss functions. The optimisation of each type of DL model architecture falls outside the scope of this paper.
As a DA method, the proposed methodology offers tremendous advantages from the perspective of fault detection as follows:
  • The latent features of the normal and faulty source samples can be distinguished because the anomaly detector is used to train the feature extractor using normal and faulty source data, which is beneficial for a subsequent one-sided DA.
  • The target domain distribution is aligned toward the source domain distribution on the latent space in a one-sided way, using the anomaly detector built in Step 1.
  • The one-sided DA using the anomaly detector enables normal and faulty target samples to be distinctive, and to be detected by the same anomaly detector trained on a rich source domain dataset.
  • Our approach ensures that the latent source and target features are identical to each other by employing the adversarial training process and a few faulty samples.
The identification of three key research gaps is detailed in Section 1. The resolution of the first and second gaps is achieved through the first, second, and third advantages of our proposed method. The fourth advantage specifically addresses the challenges presented by the third research gap. A quantitative validation of how these advantages effectively address the respective gaps is provided in Section 3. Therefore, the advantages described above enable fault detection across different domains with high reliability and bridge the gap between theoretical domain adaptation research and tangible industrial applications.

2.3. Training and Test Datasets

The training and test datasets are summarised in Table 3. The training dataset was used to build the feature extractor, feature generator, and anomaly detector, as described in Section 2.2; the test dataset was then employed to validate the proposed DA method. The training and test samples were selected randomly from the dataset. It is notable that only 10 normal and 5 faulty target samples were used to train the models.

2.4. Validation Performance Metrics

A confusion matrix can be used to evaluate the effectiveness of a fault detection system. This is a two-dimensional table containing the frequencies at which samples in each category are accurately identified or incorrectly labelled as belonging to another category. For binary classification in fault detection, the confusion matrix represents four scenarios: positive (faulty) cases can be either correctly identified or missed, and negative (normal) cases may be accurately identified or missed. These outcomes are characterised as true positive (TP), false negative (FN), true negative (TN), and false positive (FP) rates. These rates make up the confusion matrix, which is then used to calculate three performance indicators that are widely used in the industrial sector [40], as given in the following equations:
P r e c i s i o n   P = T P T P + F P
R e c a l l   R = T P T P + F N
F 1   s c o r e = 2 P R P + R
In general, the precision quantifies the proportion of samples correctly predicted as positive, while the recall represents the extent to which positive predictions correctly capture positive samples. Optimising precision and recall involves a trade-off [41]; for example, perfect recall can be achieved by predicting all samples as positive, but this results in very low precision due to numerous false alarms. In contrast, the precision will be perfect if a model predicts only the most likely positive sample as positive and the rest as negative, but this approach will result in a very low recall. One method of considering both precision and recall simultaneously is to compute their harmonic mean, referred to as the F1 score, as given in Equation (18). In this research, the F1 score, which varies from zero to one, is used to assess the fault detection accuracy. A higher F1 score indicates greater accuracy in detecting faults; the opposite is true for a lower score.

2.5. Alternative DA Models for Comparison Purposes

In this study, other types of DA models were built for comparison purposes. The following approaches were implemented:
(1)
Transfer component analysis (TCA): The primary goal of this approach is to search the feature subspace of different domains (or the source and target domains), where the domain shift between them is minimised [42]. The TCA algorithm tries to learn some transfer components across domains in an RKHS. Within the subspace defined by these transfer components, the characteristics of the data are preserved, and the data distributions across various domains are closely aligned. Logistic regression is selected as the classification method.
(2)
DANN: The aim of this approach is to find a new representation of the input features in which the source and target data cannot be distinguished by any discriminator network [28]. This new representation is learned by an encoder network in an adversarial fashion. A task network is trained on the encoded space in parallel to the encoder and discriminator networks.
These models are well-known DA methods, and have been used in many previous research papers as benchmark models. In order to build these two models, a publicly available library called the Awesome Domain Adaptation Python Toolbox (ADAPT) [43] was used. A labelled source dataset and an unlabelled target dataset were used for training, as these two models are based on an unsupervised DA method. In contrast, labelled source and target datasets, which included a few faulty samples, were used to build the proposed model, as shown in Table 3. It could be argued that comparing the proposed model with unsupervised DA methods is unfair; however, to the best of our knowledge, there is no representative DA model similar to ours. In order to make the comparative study fairer, the number of available target samples were increased, as shown in Table 4, compared to the dataset for the proposed method, shown in Table 3. The cross-validation, which typically relies on labelled data for evaluating model performance, is not applicable as target training data is unlabelled, as given in Table 4. Therefore, the randomly selected training and test samples were used, as is explained in Section 2.3.
A third model was also built, as follows:
(3)
DL model from scratch: This model was trained using only the target training dataset summarised in Table 3, in which there were 10 normal and five faulty samples. Figure 6 and Table 5 show the network architecture and hyperparameters for the model, which was previously used as a comparative model in [44].

2.6. Sensitivity Analysis

Sensitivity analysis is a common practice in technological fields, and is carried out to examine how variations in model parameters affect the output of a model [45]. To explore the sensitivity of the model to the probability threshold, a receiver operating characteristic (ROC) curve is employed. A ROC gives a comprehensive overview of the trade-off between the false positive rate (FPR) and true positive rate (TPR). The optimal ROC curve has an FPR of zero and a TPR of one. A metric known as the area under the ROC Curve (AUC) is determined by calculating the area under the complete ROC curve between (0, 0) and (1, 1). The AUC offers a single-value overview of the ROC curve, and has a value of one in the case of an ideal ROC curve. In this research, the ROC curve was plotted to correspond to the probability threshold to detect faulty samples. The probability was determined using the sigmoid function of the anomaly detector, as given in Table 1.

3. Results and Discussion

3.1. Performance Metrics and Sensitivity Analysis

The performance metrics and confusion matrices are given in Table 6 and Figure 7. The highest fault detection accuracy was achieved by the proposed method, with an F1 score of 97.9%, while DANN and the DL model from scratch had considerably lower fault detection accuracies, with F1 scores of 66.6% and 73.4%, respectively. The precision and F1 score for TCA could not be calculated, as all of the predictions were normal, as shown in Figure 7. The results reveal that a fault detection model for the target domain, which in this paper was a railway door system, can be built accurately by our DA method using source domain data from a linear actuator test rig dataset.
However, it is also necessary to ensure that the fault detection model does not have a high sensitivity to the probability threshold that is used to determine whether or not a sample is faulty. A sensitivity analysis was performed only for the proposed DA method and the DL model from scratch, as there was no need for a sensitivity analysis of TCA and DANN in view of the low fault detection accuracy (Table 6). As shown in Figure 8 and Table 7, the ROC curve for the proposed DA model was much closer to the ideal ROC curve than that of the DL model from scratch. The values of AUC for the proposed DA and DL models from scratch were 0.9976 and 0.8484, respectively. The ROC and AUC results indicate that the proposed DA method was the most accurate and the least sensitive to the threshold, and hence the most robust against variation in the input features. The proposed DA model was therefore the most accurate and robust of all the alternative models.

3.2. t-SNE Visualisation

The feature distributions of the test dataset were visualised with t-SNE, as shown in Figure 9. Four distinct distributions (relating to the source normal, source faulty, target normal, and target faulty samples) can be found in the original feature space, as illustrated in Figure 9, due to the domain shift. Suppose a fault detection model is built based on traditional ML and DL by using source domain data. In that case, fault detection accuracy should be considerably degraded when the model is applied to target domain data due to domain shift. The assumption underlying traditional ML and DL approaches is that the distribution of the test data is identical to that of the training data; however, these distributions may differ due to the different components and detailed specifications of the machinery, as clearly seen in Figure 9. In addition, traditional ML and DL typically do not possess DA capabilities. In contrast, the latent features of the normal and faulty target distributions were aligned toward the normal and faulty source distributions in the proposed method, as can be seen from Figure 9. A significant finding was that the aligned normal and faulty samples were sufficiently distinctive to be classified by the fault detection model. This clear distinction between the normal and faulty samples can be attributed to the anomaly detector that was employed to separate the two distributions, while the feature generator was trained to align the source and target data. Thus, the anomaly detector can detect faulty target samples with a high level of accuracy, as shown by the qualitative performance validation in Table 6.
However, although each feature distribution of TCA could be closer than the original features, these were unaligned, as shown in Figure 9. The poor fault detection accuracy of TCA, as shown in Table 6, may be due to this misalignment between the two domains. The misalignment of TCA and DANN may also be related to the level of similarity between the source and target domain, which might be insufficient for these models. The low similarity can be attributed to the different components involved, and specifically to railway door systems and the linear actuator test rig. These methods are therefore incapable of correctly adapting the source and target domains to become identical, despite being successful candidates as representative DA methods. Thus, the proposed DA method enables alignment between two distributions as well as clear separation between normal and faulty samples, even though the similarity is relatively low, whereas other models are unable to align the two.

3.3. Limitations

It should be emphasised that a small number of target faulty samples need to be employed with the proposed method, which means that our methodology is not unsupervised DA. In addition, only one application was used to validate the performance of the model, and further validation may be needed in the future. However, this research shows that the two domains can be aligned even when the level of similarity is low and only a few faulty samples in the target domain are used. Thus, our method is reliable and applicable to real-world industrial settings.

4. Conclusions

A novel fault detection approach based on one-sided DA using GANs for railway door systems has been proposed. In this study, the source and target domain data were drawn from a linear actuator test rig and a real-world railway door, respectively. Firstly, the anomaly detector and feature generator were trained using the label-rich source domain data to generate distinctive source latent features, and the target domain data were then aligned with the latent source features in a one-sided way. To the best of our knowledge, this is the first paper to propose a fault detection approach based on DA for railway door systems.
As a result, the performance metrics and sensitivity analysis results showed that the proposed method is the most accurate, with an F1 score of 97.9%, and is also the most robust against variation in the input features. Thus, the proposed method enables faulty target samples to be aligned with the source samples and detected accurately by the same anomaly detector, which is built with rich source data. This results in high reliability of fault detection for real-world applications despite the low level of similarity between different domains. Hence, the proposed method is not only the most accurate and robust compared to alternative models but also bridges the gap between theoretical domain adaptation research and tangible industrial applications.
In future research, it would be valuable to quantify the similarity between domains in order to be able to apply DA methods while maintaining high reliability. This is because the degree of similarity between the two domains that is required in order to be able to apply the DA method successfully is unclear. The method proposed in this research used a few faulty samples from a target domain to tackle this issue; however, even a few faulty samples from a target domain may sometimes be unavailable. An unsupervised DA method would then need to be employed, in which case the required degree of similarity between the two domains would be unknown. Addressing these issues could represent a direction for future work.

Author Contributions

Conceptualization, M.S., Y.Z. and N.P.A.; methodology, M.S.; software, M.S.; validation, M.S.; formal analysis, M.S.; investigation, M.S.; resources, M.S.; data curation, M.S.; writing—original draft preparation, M.S.; writing—review and editing, Y.Z.; visualization, M.S.; supervision, Y.Z. and N.P.A.; project administration, N.P.A.; funding acquisition, N.P.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Publicly available datasets were analyzed in this study. This data can be found here: https://figshare.com/s/dac98f9c1bc46b7a2800 (accessed on 2 November 2023).

Acknowledgments

The authors wish to thank Unipart Rail and Instrumentel for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. BSI Standards BS ISO 13372:2012; Condition Monitoring and Diagnostics of Machines—Vocabulary. BSI Standards Publication: London, UK, 2012.
  2. Cauffriez, L.; Loslever, P.; Caouder, N.; Turgis, F.; Copin, R. Robustness Study and Reliability Growth Based on Exploratory Design of Experiments and Statistical Analysis: A Case Study Using a Train Door Test Bench. Int. J. Adv. Manuf. Technol. 2012, 66, 27–44. [Google Scholar] [CrossRef]
  3. Yang, Y.; Fu, P.; He, Y. Bearing Fault Automatic Classification Based on Deep Learning. IEEE Access 2018, 6, 71540–71554. [Google Scholar] [CrossRef]
  4. Chen, Z.Q.; Li, C.; Sanchez, R.V. Gearbox Fault Identification and Classification with Convolutional Neural Networks. Shock Vib. 2015, 2015, 390134. [Google Scholar] [CrossRef]
  5. Kim, S.; Choi, J.-H. Convolutional Neural Network for Gear Fault Diagnosis Based on Signal Segmentation Approach. Struct. Health Monit. 2019, 18, 1401–1415. [Google Scholar] [CrossRef]
  6. Wang, J.; Zhao, R.; Wang, D.; Yan, R.; Mao, K.; Shen, F. Machine Health Monitoring Using Local Feature-Based Gated Recurrent Unit Networks. IEEE Trans. Ind. Electron. 2017, 65, 1539–1548. [Google Scholar] [CrossRef]
  7. Zhao, R.; Yan, R.; Wang, J.; Mao, K. Learning to Monitor Machine Health with Convolutional Bi-Directional LSTM Networks. Sensors 2017, 17, 273. [Google Scholar] [CrossRef] [PubMed]
  8. Lu, C.; Wang, Z.Y.; Qin, W.L.; Ma, J. Fault Diagnosis of Rotary Machinery Components Using a Stacked Denoising Autoencoder-Based Health State Identification. Signal Process. 2017, 130, 377–388. [Google Scholar] [CrossRef]
  9. Yan, W.; Yu, L. On Accurate and Reliable Anomaly Detection for Gas Turbine Combustors: A Deep Learning Approach. In Proceedings of the Annual Conference of the Prognostics and Health Management Society, PHM, Scottsdale, AZ, USA, 21–26 September 2019; pp. 440–447. [Google Scholar] [CrossRef]
  10. Chen, Z.; Li, W. Multisensor Feature Fusion for Bearing Fault Diagnosis Using Sparse Autoencoder and Deep Belief Network. IEEE Trans. Instrum. Meas. 2017, 66, 1693–1702. [Google Scholar] [CrossRef]
  11. Yoon, A.S.; Lee, T.; Lim, Y.; Jung, D.; Kang, P.; Kim, D.; Park, K.; Choi, Y. Semi-Supervised Learning with Deep Generative Models for Asset Failure Prediction. arXiv 2017, arXiv:1709.00845. [Google Scholar] [CrossRef]
  12. Tao, J.; Liu, Y.; Yang, D. Bearing Fault Diagnosis Based on Deep Belief Network and Multisensor Information Fusion. Shock Vib. 2016, 2016, 9306205. [Google Scholar] [CrossRef]
  13. Ben-David, S.; Blitzer, J.; Crammer, K.; Kulesza, A.; Pereira, F.; Vaughan, J.W. A Theory of Learning from Different Domains. Mach. Learn 2010, 79, 151–175. [Google Scholar] [CrossRef]
  14. Zhao, Z.; Zhang, Q.; Yu, X.; Sun, C.; Wang, S.; Yan, R.; Chen, X. Applications of Unsupervised Deep Transfer Learning to Intelligent Fault Diagnosis: A Survey and Comparative Study. IEEE Trans. Instrum. Meas. 2021, 70, 1–28. [Google Scholar] [CrossRef]
  15. Zhang, R.; Tao, H.; Wu, L.; Guan, Y. Transfer Learning With Neural Networks for Bearing Fault Diagnosis in Changing Working Conditions. IEEE Access 2017, 5, 14347–14357. [Google Scholar] [CrossRef]
  16. Zhang, C.; Xu, L.; Li, X.; Wang, H. A Method of Fault Diagnosis for Rotary Equipment Based on Deep Learning. In Proceedings of the 2018 Prognostics and System Health Management Conference, PHM-Chongqing 2018, Chongqing, China, 26–28 October 2018; IEEE: Piscataway, NJ, USA, 2019; pp. 958–962. [Google Scholar] [CrossRef]
  17. Zhang, W.; Li, C.; Peng, G.; Chen, Y.; Zhang, Z. A Deep Convolutional Neural Network with New Training Methods for Bearing Fault Diagnosis under Noisy Environment and Different Working Load. Mech. Syst. Signal Process. 2018, 100, 439–453. [Google Scholar] [CrossRef]
  18. Li, Y.; Wang, N.; Shi, J.; Liu, J.; Hou, X. Revisiting Batch Normalization For Practical Domain Adaptation. In Proceedings of the 5th International Conference on Learning Representations, ICLR 2017—Workshop Track Proceedings, Toulon, France, 24–26 April 2017. [Google Scholar] [CrossRef]
  19. Zhang, W.; Peng, G.; Li, C.; Chen, Y.; Zhang, Z. A New Deep Learning Model for Fault Diagnosis with Good Anti-Noise and Domain Adaptation Ability on Raw Vibration Signals. Sensors 2017, 17, 425. [Google Scholar] [CrossRef] [PubMed]
  20. Kim, T.; Lee, S. A Novel Unsupervised Clustering and Domain Adaptation Framework for Rotating Machinery Fault Diagnosis. IEEE Trans. Ind. Inf. 2023, 19, 9404–9412. [Google Scholar] [CrossRef]
  21. Sun, B.; Saenko, K. Deep CORAL: Correlation Alignment for Deep Domain Adaptation. In Proceedings of the Computer Vision–ECCV 2016 Workshops, Amsterdam, The Netherlands, 8–10, 15–16 October 2016; Part III 14; LNCS. Springer: Cham, Switzerland, 2016; Volume 9915, pp. 443–450. [Google Scholar] [CrossRef]
  22. Borgwardt, K.M.; Gretton, A.; Rasch, M.J.; Kriegel, H.P.; Schölkopf, B.; Smola, A.J. Integrating Structured Biological Data by Kernel Maximum Mean Discrepancy. Bioinformatics 2006, 22, e49–e57. [Google Scholar] [CrossRef]
  23. Guo, L.; Lei, Y.; Xing, S.; Yan, T.; Li, N. Deep Convolutional Transfer Learning Network: A New Method for Intelligent Fault Diagnosis of Machines with Unlabeled Data. IEEE Trans. Ind. Electron. 2019, 66, 7316–7325. [Google Scholar] [CrossRef]
  24. Long, M.; Cao, Y.; Wang, J.; Jordan, M.I.; Edu, J. Learning Transferable Features with Deep Adaptation Networks. Proc. Mach. Learn. Res. 2015, 37, 97–105. [Google Scholar]
  25. Li, X.; Zhang, W.; Ding, Q.; Sun, J.-Q. Multi-Layer Domain Adaptation Method for Rolling Bearing Fault Diagnosis. Signal Process. 2019, 157, 180–197. [Google Scholar] [CrossRef]
  26. Long, M.; Wang, J.; Ding, G.; Sun, J.; Yu, P.S. Transfer Feature Learning with Joint Distribution Adaptation. In Proceedings of the 2013 IEEE International Conference on Computer Vision, Sydney, NSW, Australia, 1–8 December 2013; pp. 2200–2207. [Google Scholar] [CrossRef]
  27. Goodfellow, I.J.; Pouget-Abadie, J.; Mirza, M.; Xu, B.; Warde-Farley, D.; Ozair, S.; Courville, A.; Bengio, Y. Generative Adversarial Nets. Adv. Neural Inf. Process Syst. 2014, 3, 2672–2680. [Google Scholar] [CrossRef]
  28. Ganin, Y.; Ustinova, E.; Ajakan, H.; Germain, P.; Larochelle, H.; Laviolette, F.; Marchand, M.; Lempitsky, V. Domain-Adversarial Training of Neural Networks. J. Mach. Learn. Res. 2016, 17, 1–35. [Google Scholar]
  29. Wang, Q.; Michau, G.; Fink, O. Domain Adaptive Transfer Learning for Fault Diagnosis. In Proceedings of the 2019 Prognostics and System Health Management Conference (PHM-Paris), Paris, France, 2–5 May 2019; IEEE: Piscataway, NJ, USA, 2019; pp. 279–285. [Google Scholar]
  30. Chen, Z.; He, G.; Li, J.; Liao, Y.; Gryllias, K.; Li, W. Domain Adversarial Transfer Network for Cross-Domain Fault Diagnosis of Rotary Machinery. IEEE Trans. Instrum. Meas. 2020, 69, 8702–8712. [Google Scholar] [CrossRef]
  31. Cheng, C.; Zhou, B.; Ma, G.; Wu, D.; Yuan, Y. Wasserstein Distance Based Deep Adversarial Transfer Learning for Intelligent Fault Diagnosis with Unlabeled or Insufficient Labeled Data. Neurocomputing 2020, 409, 35–45. [Google Scholar] [CrossRef]
  32. Wang, Q.; Michau, G.; Fink, O. Missing-Class-Robust Domain Adaptation by Unilateral Alignment. IEEE Trans. Ind. Electron. 2021, 68, 663–671. [Google Scholar] [CrossRef]
  33. Saito, K.; Watanabe, K.; Ushiku, Y.; Harada, T. Maximum Classifier Discrepancy for Unsupervised Domain Adaptation. In Proceedings of the IEEE Computer Society Conference on Computer Vision and Pattern Recognition, Honolulu, HI, USA, 21–26 July 2017; pp. 3723–3732. [Google Scholar] [CrossRef]
  34. Ruiz-Carcel, C.; Starr, A. Data-Based Detection and Diagnosis of Faults in Linear Actuators. IEEE Trans. Instrum. Meas. 2018, 67, 2035–2047. [Google Scholar] [CrossRef]
  35. López de Calle-Etxabe, K.; Ruiz-Cárcel, C.; Starr, A.; Ferreiro, S.; Arnaiz, A.; Gómez-Omella, M.; Sierra, B. Hybrid Modelling for Linear Actuator Diagnosis in Absence of Faulty Data Records. Comput Ind 2020, 123, 103339. [Google Scholar] [CrossRef]
  36. Ruiz-Carcel, C.; Starr, A. Data Set for “Data-Based Detection and Diagnosis of Faults in Linear Actuators”. Available online: https://figshare.com/s/dac98f9c1bc46b7a2800 (accessed on 22 July 2022).
  37. Shimizu, M.; Perinpanayagam, S.; Namoano, B. A Real-Time Fault Detection Framework Based on Unsupervised Deep Learning for Prognostics and Health Management of Railway Assets. IEEE Access 2022, 10, 96442–96458. [Google Scholar] [CrossRef]
  38. Namoano, B. Fault Diagnosis in Time Series Data with Application to Railway Assets. Ph.D. Thesis, School of Aerospace, Transport and Manufacturing, Engineering And Management of Manufacturing Systems, Cranfield University, Cranfield, UK, 2020. [Google Scholar]
  39. Gretton, A.; Borgwardt, K.M.; Rasch, M.J.; Smola, A.; Schölkopf, B.; Smola Gretton, A. A Kernel Two-Sample Test. J. Mach. Learn. Res. 2012, 13, 723–773. [Google Scholar]
  40. Jennions Ian, K. Integrated Vehicle Health Management: The Technology; IEEE: Piscataway, NJ, USA, 2013. [Google Scholar]
  41. Andreas, C.; Müller, S.G. 5. Model Evaluation and Improvement. In Introduction to Machine Learning with Python; O’Reilly Media, Inc.: Sebastopol, CA, USA, 2016. [Google Scholar]
  42. Pan, S.J.; Tsang, I.W.; Kwok, J.T.; Yang, Q. Domain Adaptation via Transfer Component Analysis. IEEE Trans. Neural. Netw. 2011, 22, 199–210. [Google Scholar] [CrossRef]
  43. De Mathelin, A.; Atiq, M.; Richard, G.; De La Concha, A.; Yachouti, M.; Deheeger, F.; Mougeot, M.; Vayatis, N. ADAPT: Awesome Domain Adaptation Python Toolbox. arXiv 2021, arXiv:2107.03049. [Google Scholar]
  44. Shimizu, M.; Perinpanayagam, S.; Namoano, B. A Fault Detection Technique Based on Deep Transfer Learning from Experimental Linear Actuator to Real-World Railway Door Systems. In Proceedings of the Annual Conference of the Prognostics and Health Management Society (Accepted), Turin, Italy, 6–8 July 2022. [Google Scholar]
  45. Russell, S.; Norvig, P. Artificial Intelligence: A Modern Approach; Pearson Education, Limited: London, UK, 2021; ISBN 9781292401171. [Google Scholar]
Figure 1. Three-dimensional model of the test rig and lateral view of the rig [36].
Figure 1. Three-dimensional model of the test rig and lateral view of the rig [36].
Sensors 23 09688 g001
Figure 2. Examples of position measurements and current signals from the linear actuator test rig.
Figure 2. Examples of position measurements and current signals from the linear actuator test rig.
Sensors 23 09688 g002
Figure 3. Normal and faulty current signals from the linear actuator test rig.
Figure 3. Normal and faulty current signals from the linear actuator test rig.
Sensors 23 09688 g003
Figure 4. Current signals for door systems, and normal and faulty signals for the closing operation.
Figure 4. Current signals for door systems, and normal and faulty signals for the closing operation.
Sensors 23 09688 g004
Figure 5. Workflow for the proposed methodology.
Figure 5. Workflow for the proposed methodology.
Sensors 23 09688 g005
Figure 6. DL model from scratch [44].
Figure 6. DL model from scratch [44].
Sensors 23 09688 g006
Figure 7. Confusion matrix.
Figure 7. Confusion matrix.
Sensors 23 09688 g007
Figure 8. ROC curves.
Figure 8. ROC curves.
Sensors 23 09688 g008
Figure 9. t-SNE visualisation: (i) original features; (ii) latent features for the proposed method; (iii) latent features for TCA.
Figure 9. t-SNE visualisation: (i) original features; (ii) latent features for the proposed method; (iii) latent features for TCA.
Sensors 23 09688 g009
Table 1. Network architecture and numbers of learnable parameters.
Table 1. Network architecture and numbers of learnable parameters.
ModelLayerOutput ShapeParameters
Feature extractorLinear1 × 10020,000
ReLU1 × 100-
Linear1 × 505050
ReLU1 × 500.0001
Linear1 × 10510
Anomaly detectorLinear1 × 555
ReLU1 × 5-
Linear1 × 16
Sigmoid1 × 1-
Feature generatorLinear1 × 10020,000
ReLU1 × 100-
Linear1 × 505050
ReLU1 × 500.0001
Linear1 × 10510
DiscriminatorLinear1 × 333
ReLU1 × 3-
Linear1 × 14
Sigmoid1 × 1-
Table 2. Hyperparameters for the models.
Table 2. Hyperparameters for the models.
ModelHyperparameter NameHyperparameter
Feature extractorOptimiserAdam
Learning rate0.001
Max epoch5000
Anomaly detectorOptimiserAdam
Learning rate0.001
Max epoch5000
Feature generatorOptimiserAdam
Learning rate0.001
Max epoch8000
Feature generatorOptimiserAdam
Learning rate0.1
Max epoch8000
Table 3. Training and test datasets.
Table 3. Training and test datasets.
Training/TestDomainNormalFaultyTotal
TrainingSource5050100
TrainingTarget10515
TestTarget5050100
Table 4. Training and test datasets for comparison models.
Table 4. Training and test datasets for comparison models.
Training/TestDomainLabelNormalFaultyTotal
TrainingSourceLabelled5050100
TrainingTargetUnlabelled100100
TestTargetLabelled5050100
Table 5. Hyperparameters of a fault detection model for the DL model from scratch [44].
Table 5. Hyperparameters of a fault detection model for the DL model from scratch [44].
LayerHyperparameter NameHyperparameter
Whole layersOptimiserAdam
Max epoch3000
Mini-batch size120
Learning rate0.0001
LSTMActivation function for the hidden stateTanh
Activation function for the gatesSigmoid
Number of activation units40
FCNumber of hidden units2
Table 6. Fault detection accuracy.
Table 6. Fault detection accuracy.
ModelPrecision (%)Recall (%)F1 Score (%)
Proposed DA model1009697.9
TCANaN50NaN
DANN5010066.6
DL model from scratch75.072.073.4
Table 7. AUC values.
Table 7. AUC values.
ModelProposed MethodDL Model from Scratch
AUC0.99760.8484
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.

Share and Cite

MDPI and ACS Style

Shimizu, M.; Zhao, Y.; Avdelidis, N.P. A Fault Detection Approach Based on One-Sided Domain Adaptation and Generative Adversarial Networks for Railway Door Systems. Sensors 2023, 23, 9688. https://doi.org/10.3390/s23249688

AMA Style

Shimizu M, Zhao Y, Avdelidis NP. A Fault Detection Approach Based on One-Sided Domain Adaptation and Generative Adversarial Networks for Railway Door Systems. Sensors. 2023; 23(24):9688. https://doi.org/10.3390/s23249688

Chicago/Turabian Style

Shimizu, Minoru, Yifan Zhao, and Nicolas P. Avdelidis. 2023. "A Fault Detection Approach Based on One-Sided Domain Adaptation and Generative Adversarial Networks for Railway Door Systems" Sensors 23, no. 24: 9688. https://doi.org/10.3390/s23249688

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop