Next Article in Journal
ID-Det: Insulator Burst Defect Detection from UAV Inspection Imagery of Power Transmission Facilities
Previous Article in Journal
UAV Imagery-Based Classification Model for Atypical Traditional Village Landscapes and Their Spatial Distribution Pattern
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Deep Learning for Indoor Pedestal Fan Blade Inspection: Utilizing Low-Cost Autonomous Drones in an Educational Setting

by
Angel A. Rodriguez
,
Mason Davis
,
Joshua Zander
,
Edwin Nazario Dejesus
,
Mohammad Shekaramiz
*,
Majid Memari
and
Mohammad A. S. Masoum
Machine Learning and Drone Laboratory, Electrical and Computer Engineering Department, Utah Valley University, Orem, UT 84058, USA
*
Author to whom correspondence should be addressed.
Drones 2024, 8(7), 298; https://doi.org/10.3390/drones8070298
Submission received: 6 June 2024 / Revised: 28 June 2024 / Accepted: 30 June 2024 / Published: 5 July 2024

Abstract

:
This paper introduces a drone-based surrogate project aimed at serving as a preliminary educational platform for undergraduate students in the Electrical and Computer Engineering (ECE) fields. Utilizing small Unmanned Aerial Vehicles (sUAVs), this project serves as a surrogate for the inspection of wind turbines using scaled-down pedestal fans to replace actual turbines. This approach significantly reduces the costs, risks, and logistical complexities, enabling feasible and safe on-campus experiments. Through this project, students engage in hands-on applications of Python programming, computer vision, and machine learning algorithms to detect and classify simulated defects in pedestal fan blade (PFB) images. The primary educational objectives are to equip students with foundational skills in autonomous systems and data analysis, critical for their progression to larger scale projects involving professional drones and actual wind turbines in wind farm settings. This surrogate setup not only provides practical experience in a controlled learning environment, but also prepares students for real-world challenges in renewable energy technologies, emphasizing the transition from theoretical knowledge to practical skills.

1. Introduction

The emergence of drone technology coupled with advances in machine learning and computer vision has paved the way for innovative educational approaches in the engineering and technology disciplines. Drones, or Unmanned Aerial Vehicles (UAVs), have revolutionized various industries, including agriculture, surveillance, logistics, and environmental monitoring. Their ability to capture high-resolution images and data from inaccessible areas has made them indispensable tools in modern technology. Moreover, the integration of computer vision with drone technology has significantly enhanced autonomous navigation, object detection, and environmental mapping capabilities.
This paper builds upon the foundation of our indoor vision-based path-planning surrogate project and investigates the necessary changes required for blade inspection and automated analysis of blade health [1]. The project aims to simulate real-world wind turbine inspections using scaled-down pedestal fans as stand-ins for actual wind turbines. This approach reduces the operational risks and costs, making it an ideal preliminary research project for educational purposes. By employing scaled-down models, students can engage in practical, hands-on learning without the logistical and safety challenges associated with full-scale wind turbines.
The educational initiative described in this paper introduces undergraduate students to the practical applications of autonomous systems. By bridging the gap between theoretical concepts taught in classrooms and real-world technological implementations, students gain valuable insights into the complexities and challenges of modern engineering problems. The project not only fosters an understanding of drone navigation and computer vision techniques, but also prepares students for subsequent large-scale projects involving professional drones and real wind turbines in wind farms.
The contributions of this work are as follows:
  • Deep learning classification of blade faults: Our investigation of deep learning classification algorithms, including AlexNet, VGG, Xception, ResNet, DenseNet, and ViT, presents an introduction to key novel architectures along with their contributions to the field of computer vision. This exploration lays the foundational history of deep learning, enabling those new to the field to identify architectural design patterns and key mathematical characteristics vital to visual classification.
  • Path planning for blade inspection: Real-time image detection and decision-making enhance drone adaptability, allowing for seamless inspection of pedestal fan blades (PFBs). This component demonstrates how drones can be programmed to navigate complex environments autonomously, avoiding obstacles and optimizing inspection routes.
  • Educational aspects and applications: Through detailed explanations, publicly available code, and low-cost hardware requirements, this project offers practical experience in developing drone platforms infused with machine learning and computer vision. This provides students with a real-world example to enhance comprehension and engagement with classroom materials in machine learning, controls, and autonomy. The educational impact is significant as it prepares students for careers in emerging technologies.
The remainder of this paper is organized as follows: Section 2 provides an overview of the proposed integration into the machine learning course and the development of a new course on drone technology. This is followed by a brief literature review of existing work in this area, with gaps identified. Section 4 presents the methodology of this study including the inspection method utilized for data generation and the deep learning classification architecture overviews. This is followed by the resulting classification performance on the PFBs dataset for each architecture in our study and a discussion on the implications and findings. Section 6 provides a discussion on the real-world challenges and alternative data acquisition strategies. Finally, conclusions are drawn, followed by documentation of publicly available material related to the project.

2. A Hands-On Project for Machine Learning and Autonomous Drones

This research project was conducted in part as a final project for the existing Machine Learning (ECE 4850) course at the Electrical and Computer Engineering Department of Utah Valley University in Spring 2024. Based on the instructor’s evaluation and the students’ feedback at the end of the semester, the students in this course appeared to gain valuable knowledge on the autonomy of drones integrated with machine learning and computer vision. This course of Machine Learning is fully project-based, where the students are assigned five bi-weekly projects and one final project related to various subjects of machine learning, deep learning, probabilistic modeling, and real-time object detection via a low-cost drone. In the current structure, a course in Python and one in probability and statistics are the prerequisites for this course. This allows students from the Electrical Engineering, Computer Engineering, and Computer Science majors to be able to register for this course in their Junior or Senior year. In this course, the students will be exposed to several machine learning algorithms including supervised and unsupervised learning, the fundamentals of neural networks, and recent advancements and architectures in deep learning. The bi-weekly projects tend to gauge their understanding of the concepts and challenge their critical thinking to come up with solutions to solve real-world problems. For their final project, the students will work on the proposed project in this paper and its complement published in [1]. Three lectures are dedicated to teaching the students how to work and program the Tello EDU drone using Python with a complement of 12 short video series that are currently under development.
Moving forward, this applied project will be integrated into the content of a new course in AI and Smart Systems (AISS) that is under development as an elective course. The course of AISS will contain materials related to both machine/deep learning and autonomous drones. The pre-requisite for this course will be a course on Python programming and Machine Learning. Thus, the students taking this course will be already familiar with the main concepts of machine/deep learning algorithms and architectures. In the AISS course, the students will become familiarized with drone programming through Python and Kotlin, several path-planning algorithms such as A * , Rapidly-exploring Random Tree ( R R T ), Fast R R T , and R R T * , as well as the implementation of machine/learning techniques for classification and object detection. The course will be fully project-based, and the students will work in groups of two.
This project is designed to provide undergraduate students with a robust foundation in both the theoretical and practical aspects of machine learning and autonomous systems, specifically focusing on drone technology. The primary objectives of this project are to introduce students to the principles of drone technology and autonomous systems, equip them with practical skills in programming, path planning, machine/deep learning, image processing, and computer vision, and provide them with a good understanding of the real-world applications and challenges in the field of drones. This project leverages computer vision and deep learning techniques to analyze and interpret visual data captured by drones, which is crucial for detecting and classifying objects and performing anomaly detection.
Path planning and image processing, leveraging machine learning and deep learning, are integral components of this project due to their pivotal roles in autonomous drone operations. Path planning entails determining an optimal route for the drone, which is vital for autonomous navigation, mission optimization, and practical applications such as search and rescue and environmental monitoring. Computer vision allows the analysis of visual data captured by the drone, which is essential for detecting and classifying objects, performing anomaly detection, and integrating AI to enhance the comprehension of autonomous systems. These technologies are deeply interconnected. For example, real-time computer vision is crucial for identifying and avoiding obstacles, a fundamental aspect of effective path planning. In target-tracking applications, computer vision supplies essential data for path-planning algorithms to dynamically adjust the drone’s trajectory. Furthermore, computer vision facilitates the creation of detailed environmental maps, which path-planning algorithms utilize to optimize navigation routes.
Students undertake various technical tasks throughout this project, including programming drones to control the behavior and maneuvers, developing and implementing path-planning and computer vision algorithms, and designing and conducting experiments to test drone capabilities. These tasks are designed to enhance problem-solving skills, technical proficiency, critical thinking, and collaboration and communication abilities. Through these activities, students gain hands-on experience with cutting-edge technologies, analyze and interpret data to make informed decisions, and work in teams to effectively present their findings. This comprehensive approach ensures that students are well prepared for careers in the field of drones and autonomous systems.
Optimizing the path planning for drone inspections of PFBs unlocks numerous advantages, crucially enhancing the inspection process. Key among these is achieving expansive inspection coverage, where path-planning algorithms guide drones on optimal routes to encompass entire wind farms thoroughly. This meticulous approach ensures a comprehensive inspection, significantly mitigating the risk of overlooked turbines and minimizing redundant coverage areas. Additionally, the strategic optimization of path planning plays a pivotal role in enhancing operational efficiency. Drones can avoid unnecessary detours by navigating paths that account for distance and prevailing weather conditions, conserving flight time and battery life. This efficiency streamlines the inspection process and considerably reduces cost, rendering wind turbine inspections more economically feasible. With the addition of computer vision, drone adaptability increases, allowing real-time flexibility during the inspection process. This ensures drones operate securely within the complex and changing airspace of wind farms. In our previous work, simulated annealing was deployed to allow efficient paths to be generated for various combinations of PFBs; Haar-cascade classifiers were leveraged for pedestal fans’ (in replacement of wind turbines) identification and positioning; the Hough transform was utilized for further vision positioning during the inspection mission [1].
Building on this work, the investigation of deep learning for fault analysis is presented to complete the proposed surrogate project and provide a comprehensive educational approach to learning these emergent technologies. Here, the automation of wind turbine inspection is of great interest as traditional approaches are time-consuming and often dangerous. These methods include manual inspection, requiring personnel to scale turbine blades, or visual inspection using various imaging and telescopic devices. As wind turbine installations continue to scale worldwide, the need for more efficient and safe inspection is ever-present. Through imagery captured with the autonomous path-planning algorithms presented in [1], various architectures are trained and compared on the ability to determine fault status in these PFBs. Coupled with autonomous path planning, this automated classification of faults completes the inspection system and provides an educational setting to explore these deep learning models.

3. Related Works

As we embark on exploring innovative educational methods for engineering and technology, the integration of autonomous drone technology and computer vision presents a promising frontier. This section delves into significant advancements in the application of drones and computer vision for wind turbine blade (WTB) inspection. We review advanced deep learning models and path-planning algorithms, highlighting the pivotal role these technologies play in shaping current methodologies for renewable energy maintenance. This exploration situates our work within the broader scientific discourse and underscores our commitment to optimizing renewable energy maintenance through technological innovation.

3.1. UAV Navigation Techniques

Vision-based navigation has emerged as a fundamental component in UAV applications, spanning from aerial photography to intricate rescue operations and aerial refueling. The accuracy and efficiency required for these applications reinforce the critical role of computer vision and image processing across a spectrum of UAV operations. Notable challenges include establishing robust peer-to-peer connections for coordinated missions [2,3,4,5]. Innovations have also been seen in vision-based autonomous takeoff and landing that integrate IMU, GPS data, and LiDAR, although inaccuracies in attitude determination highlight the limitations of current technologies [6,7,8,9]. Furthermore, aerial imaging and inspection have benefited from advancements in UAV performance, leveraging methods for enhanced autonomy in surveillance tasks [10]. Innovations in vision-based systems offer robust alternatives for environments where traditional positioning signals are obstructed, utilizing Visual Simultaneous Localization And Mapping (VSLAM) and visual odometry to navigate through GPS-denied environments [11,12]. Extended Kalman Filter (EKF) and Particle Filter applications refine UAV positioning and orientation estimation, addressing the complexities of navigation in intricate scenarios [13,14,15]. As UAV navigation evolves, the integration of heuristic searching methods for path optimization and techniques such as the Lucas–Kanade method for optical flow demonstrate innovative approaches to altitude and speed control, significantly advancing UAV autonomy [16,17,18,19,20,21,22,23,24,25].

3.2. WTB Computer Vision-Based Inspection Methods

As the world pivots towards renewable energy, wind turbines emerge as pivotal assets, underscoring the necessity of effective maintenance strategies. Visual inspection utilizing RGB cameras has marked a significant technological leap, enhancing the efficacy and advancement of WTB inspections [26,27,28]. This exploration delves into the evolution of WTB visual inspection methodologies, accentuating the integration of data-driven frameworks, cutting-edge deep learning techniques, and drones equipped with RGB cameras for superior imaging quality [29,30,31]. Emphasis is placed on innovative Convolutional Neural Networks (CNNs) for meticulous crack detection and models like You Only Look Once (YOLO) for enhanced defect localization, evaluating the performance of various CNN architectures and anomaly-detection techniques in pinpointing surface damage on WTBs [32]. Additional research investigates model simplification to further enhance the operational efficiency and inference speed of these deep learning-powered inspection systems, allowing closer to real-time detection and deployment in computationally restrictive environments [33,34,35,36].

3.3. Educational Implementation and Surrogate Projects with Tello EDU

The Tello EDU drone, created by Ryze technology, incorporates DJI flight technology and an Intel processor allowing for a low-barrier entry into programmable drone applications. Here, the key educational benefits include an easy-to-use Python Application Programming Interface (API) for basic drone functions along with the ability to leverage abstractable block-based coding for even earlier use in programs. Here, we briefly discuss its utilization in pioneering surrogate projects that simulate real-world challenges, thereby providing students with practical experience that effectively bridges theoretical learning with hands-on application.
Pinney et al. showcased the strength of Tello EDUs in educational projects. Here, the challenges of autonomous searching, locating, and logging were investigated through the leveraging of cascading classifiers and real-time object detection. This enabled the drone to autonomously search for targets and interact with them in unique ways [37,38]. In the subject of machine learning, this drone allows the capture of aerial imagery, which can be leveraged for practical experience in data-gathering and training algorithms. Here, Seibi et al. and Seegmiller et al. approach the detection of wind turbines and their faults in unique ways. These projects allow for team-based optimization of challenges related to real-world engineering development, enhancing students’ technological literacy and fostering a deeper understanding of system integration and teamwork [39,40].
Pohudina et al. and Lane et al. used Tello EDU drones in group settings and competitive scenarios promoting an environment of active learning and innovation, demonstrating the drone’s adaptability to diverse educational needs [41,42]. Further studies by Lochtefeld et al., Zou et al., and Barhoush et al. emphasize the potential of these drones to enhance unmanned aerial operations and interactive learning through dynamic environments and responsive tasks [43,44,45]. Future educational projects are expected to delve deeper into more advanced computer vision techniques and the integration of machine learning models, broadening the scope of drone applications in education [46,47,48].
As has been shown, the Tello EDU drone serves as a pivotal educational platform that not only facilitates the exploration of cutting-edge technology, but also prepares students for the technological challenges of the future. Its continued evolution is anticipated to further revolutionize STEM education, enabling more complex, interactive, and comprehensive learning experiences.

3.4. Hardware Configuration and Alternatives

In this study, we utilized the Tello EDU drone, which is an accessible and educationally friendly platform equipped with essential features for computer vision and machine learning applications. The Tello EDU drone is configured with a 5 MP camera capable of streaming 720p video at 30 frames per second, complemented by Electronic Image Stabilization (EIS) [49]. It also features a Vision Positioning System (VPS) [50] to facilitate precise hovering and includes an Intel 14-Core Processor for high-quality footage. This drone offers programmability through the Tello SDK, allowing integration with Python scripts for automated control and image-processing tasks.
For regions where the Tello EDU is not available, several alternative drones offer similar capabilities. One such alternative is the Robolink CoDrone [51], which supports visual coding options like Blockly and additional features such as a physical controller and an LED screen. The Crazyflie 2.1 [52] is another versatile platform, suitable for advanced users with technical knowledge, offering highly customizable and programmable options. The Airblock drone [53] is modular and fun to build, making it ideal for STEM education and creative projects. Finally, the Parrot ANAFI [54] offers a high-quality 4K camera and versatile programming options and is suitable for professional applications.
To ensure the project can be replicated with different hardware, we recommend several specifications for replacement drones. The camera quality should have at least 720p resolution to ensure sufficient detail in captured images. The drones should also have programmability with access to a comprehensive SDK supporting Python or similar programming languages. Built-in WiFi operating at 2.4 GHz is necessary for connectivity to a local computer to execute drone commands and facilitate autonomous decision-making. Additionally, the drones should feature on-board processing capabilities to handle real-time image stabilization and processing tasks. For stabilization and control, a built-in IMU and sensors are essential to maintain flight stability and accurate height measurement.
The information about these drones in Table 1 may vary by country and availability. Additionally, these drones might not be easily found in all regions due to security and regulatory restrictions. Table 1 provides a detailed comparison of the different drones, highlighting their key features, programming capabilities, and additional functionalities. By carefully evaluating these factors, you can choose the drone that best aligns with your specific needs and budget. By adhering to these specifications, the project can be effectively implemented using various drones available in different regions, ensuring broader accessibility and applicability.

4. Methodology

Building on the autonomous path-planning work in [1], the visual inspection of PFBs is investigated along with deep learning classification in which the culminating inspection solution is completed. This is accomplished in two primary tasks. First, the autonomous path-planning must be adjusted to capture images of each PFB in both front and back orientations. This allows for a dataset to be built on which the second task can begin. Here, the deep learning algorithms can be trained, tested, and compared for performance in the task of PFB anomaly detection. The methodologies for each of the primary tasks are detailed below.

4.1. Drone Path Planning

To facilitate the inspection of the PFBs, the searching algorithm deployed, first discussed in our previous work [1], consists of a trained Haar-cascade classifier for PFB identification. However, to allow for precise imagery to be captured of the blades, the metal protective casings were removed here, and thus, a new model was required for this task. Additionally, the same QR code system was used to allow seamless tracking of image save locations within the application and allow for drift adjustments. Finally, the auxiliary camera present on the Tello EDU was utilized along with OpenCV’s Hough Circle transform to allow precise takeoff and landing on the designated helipad. These features of the experimental setup can be seen in Figure 1.
Providing the software framework and backbone of the entire system, a computer running the Python scripts enables seamless integration of various components including sensor data, flight control, drone communication, and multi-threaded processing for deep learning algorithms. Figure 2 illustrates the comprehensive system architecture integrating the DJI Tello EDU drone with a local computer, essential for executing the multifaceted components of our educational experiments. In this figure, computer vision-based path planning showcases how the drone utilizes real-time image data for navigation, including object recognition and environmental mapping. Cartesian coordinate-based path planning details the method by which the drone’s position is calculated in relation to a Cartesian grid, enabling precise movement based on the calculated coordinates. Trigonometric computations via object detection depict the integration of object detection with trigonometric calculations to determine the drone’s path by calculating the distances and angles relative to the detected objects. Crack and anomaly detection via deep learning illustrates the process of analyzing captured images using deep learning models to detect structural anomalies or damages such as cracks on the simulated turbine PFBs.

4.2. Overview of the Deep Learning Models Used for Classification

In this study, we focus on CNNs [55] due to their proven effectiveness in image classification tasks, particularly in computer vision applications like defect detection and object recognition. CNNs are designed to automatically and adaptively learn the spatial hierarchies of features from input images, making them highly suitable for the detailed analysis required in our application. The inherent features of CNNs, such as local connectivity, weight sharing, and the use of multiple layers, allow for the extraction of both low-level and high-level features, facilitating accurate and efficient image classification.
While CNNs have demonstrated superiority over traditional Artificial Neural Networks (ANNs) [56] in handling image data, their performance in comparison to Recurrent Neural Networks (RNNs) [57] and Vision Transformers (ViTs) [58] for the proposed application also merits discussion. RNNs, which are typically used for sequential data and time series analysis, are less effective in capturing spatial dependencies within an image. Therefore, they are not the optimal choice for our application, which requires the analysis of static images with complex spatial features.
Vision Transformers (ViTs), on the other hand, represent a novel approach in image classification, leveraging self-attention mechanisms to capture global context. Although ViTs have shown promising results, they often require larger datasets and more computational resources for training compared to CNNs. In our educational setting, where computational resources and labeled data may be limited, CNNs offer a more practical and accessible solution.
The choice of CNNs over ANNs and RNNs is also justified by their architectural advantages. ANNs, with their fully connected layers, lack the spatial hierarchy that CNNs exploit. This makes ANNs less effective for tasks involving image data, where spatial relationships between pixels are crucial. RNNs, designed for sequential data processing, excel in tasks such as natural language processing, but do not leverage the spatial structure of images effectively.
Furthermore, the historical development and success of CNN architectures, such as AlexNet, VGG, ResNet, DenseNet, and Xception, provide a robust foundation for our work. These architectures have been extensively validated across various image-classification benchmarks, and their design principles align well with the requirements of our defect-detection task. By employing CNNs, we leverage their strengths in feature extraction and pattern recognition, ensuring a reliable and efficient inspection process for pedestal fan blades.
In summary, the selection of CNNs is based on their superior ability to handle spatial hierarchies in image data, their proven success in numerous computer vision applications, and the practical considerations of computational resources and data availability in our educational setting.
The subsequent sections will discuss each of these models in detail, highlighting their architectures and the specific advantages they bring to our application.

4.2.1. AlexNet

Introduced by Krizhevsky et al., AlexNet sparked a revolution in deep learning, achieving unparalleled success in the ImageNet challenge [59]. This pioneering model, with its innovative architecture of five convolutional layers followed by three fully connected layers, demonstrated the profound capabilities of deep networks in image classification. This structure can be seen in Figure 3 with its distinct convolutional and fully connected layers. Notably, AlexNet was one of the initial models to utilize rectified linear units (ReLUs) for faster convergence and dropout layers to combat overfitting, setting new benchmarks in training deep neural networks [60]. Optimized for dual Nvidia GTX 580 GPUs, it showcased the significant advantages of GPU acceleration in deep learning. The AlexNet model can be mathematically described as follows: For each convolutional layer l, the operation can be represented as:
X ( l + 1 ) = f W ( l ) X ( l ) + b ( l )
where X ( l ) is the input feature map, W ( l ) is the filter kernel, b ( l ) is the bias term, ∗ denotes the convolution operation, and f ( · ) is the ReLU activation function defined as:
f ( x ) = max ( 0 , x )
The fully connected layers can be expressed as:
Z ( l + 1 ) = f W ( l ) Z ( l ) + b ( l )
where Z ( l ) is the input vector to the fully connected layer and W ( l ) and b ( l ) are the weight matrix and bias vector, respectively.

4.2.2. VGG

The VGG models, developed by the Visual Geometry Group at the University of Oxford, showcased the critical role of network depth in enhancing neural network performance [61]. Characterized by their use of 3 × 3 convolutional filters, VGG models efficiently detect complex features while keeping parameter counts manageable. The VGG series, including VGG11, VGG13, VGG16, and VGG19, demonstrates the benefits of increasing depth for detailed feature representation, although it raises considerations regarding overfitting and computational demands. Some VGG variants incorporate batch normalization to improve training stability. Despite their intensive computational requirements, VGG models are well known for their depth and robustness in feature extraction, proving especially beneficial in transfer learning applications [61]. The deepest network, VGG19, is visualized in Figure 4. Here, the proposed constant kernel size of 3 × 3 and stride of 1 can be seen. The VGG model’s convolutional layers can be described by the following equation:
For each convolutional layer l, the operation is:
X ( l + 1 ) = f W ( l ) * X ( l ) + b ( l )
where X ( l ) is the input feature map, W ( l ) is the filter kernel of size 3 × 3 , b ( l ) is the bias term, and f ( · ) is the ReLU activation function.
Batch normalization, when applied, can be represented as:
X ^ ( l ) = X ( l ) μ σ 2 + ϵ
Y ( l ) = γ X ^ ( l ) + β
where μ and σ 2 are the mean and variance of the input, ϵ is a small constant for numerical stability, and γ and β are learnable parameters.

4.2.3. Xception

The Inception architecture, renowned for its innovative Inception module, has significantly advanced neural network designs by enabling the capture of features at various scales. Developed with the goal of achieving both depth and width in networks without compromising computational efficiency, it introduced parallel pooling and convolutions of different sizes. Xception, or Extreme Inception, evolved from this model by employing depthwise separable convolutions, thereby enhancing the network’s ability to learn spatial features and channel correlations more effectively [62]. This technique comprises two stages: depthwise convolution and pointwise convolution.
In the first stage, depthwise convolution, a single filter is applied per input channel. Mathematically, for an input X of size W × H × D (width, height, depth), the depthwise convolution is defined as:
Y i , j , d = m = 0 k 1 n = 0 k 1 X i + m , j + n , d · F d , m , n ,
where Y i , j , d represents the output feature map, k is the filter size, F d , m , n is the filter for depth d, and ( i , j ) are the spatial coordinates in the feature map. This step involves convolving each input channel separately with a corresponding filter, preserving the depth dimension, but reducing the spatial dimensions.
In the second stage, pointwise convolution, the features obtained from the depthwise convolution are combined channelwise. This is mathematically expressed as:
Z i , j , d = d = 0 D 1 Y i , j , d · F d , d ,
where Z i , j , d is the final output after the pointwise convolution, D is the number of input channels, and F d , d is the pointwise filter. This operation effectively combines the information across different channels, allowing the model to learn more complex patterns.
The Inception module’s output, I ( x ) , incorporates various convolutional and pooling operations as follows:
I ( x ) = [ f 1 × 1 ( x ) , f 3 × 3 ( x ) , f 5 × 5 ( x ) , f p o o l ( x ) ] ,
where f 1 × 1 ( x ) , f 3 × 3 ( x ) , f 5 × 5 ( x ) , and f p o o l ( x ) represent the outputs of the 1 × 1, 3 × 3, and 5 × 5 convolutions and the pooling operations, respectively. This structure allows the network to capture features at multiple scales, contributing to its robustness and effectiveness in complex image-recognition tasks.
By refining the convolutional approaches, the Xception model demonstrates the power of deep learning techniques in achieving a more nuanced understanding of image data, enhancing both parameter efficiency and overall performance in various computer vision applications. The Xception architecture is comprised of 3 sections, the entry flow, middle flow, and exit flow, highlighted in Figure 5. Here, the entry flow leverages separable convolutions and skip connections. This is followed by the middle flow, which repeats 8 times prior to the exit flow. Finally, there are another skip connection and separable convolution with a global max pooling.

4.2.4. Residual Networks

Introduced by Kaiming He and colleagues, Residual Networks (ResNets) tackled the challenge of training deeper networks effectively [63]. At their core, ResNets employ residual blocks that aim to learn residual functions, simplifying the learning process by focusing on residual mappings. These blocks, through skip connections, allow the network to adjust identity mappings by a residual, enhancing gradient flow and addressing the vanishing gradient problem.
In a residual block, the network is tasked with learning the residual function F ( x ) instead of the original function H ( x ) . Mathematically, this is represented as:
F ( x ) = H ( x ) x ,
where F ( x ) is the residual function that the block learns, H ( x ) is the desired underlying mapping, and x is the input to the block. The output of the residual block is then given by:
Output = F ( x ) + x = H ( x ) x + x = H ( x ) ,
where Output is essentially the original function H ( x ) , with the skip connection ensuring that the input x is directly added to the learned residual function F ( x ) . This mechanism allows gradients to bypass one or more layers, thereby mitigating the vanishing gradient problem and facilitating the training of much deeper networks.
The architecture of ResNets includes various depths such as ResNet18, ResNet34, ResNet50, ResNet101, and ResNet152, each denoting the number of layers in the network. Furthermore, there are wide ResNet versions that expand the number of filters in each layer, thereby enhancing performance at the cost of increased computational demands [63]. The 50-layer deep ResNet model is visualized in Figure 6. Here, the residual connections can be seen along with the variable depth layers repeated, corresponding to the depth of the architecture chosen.
ResNets have profoundly influenced subsequent deep learning models by offering a robust solution to the challenges associated with training deep networks, balancing performance with computational efficiency. The introduction of residual blocks has set a precedent for many modern architectures, underscoring the importance of efficient gradient flow and effective training of deep models.

4.2.5. DenseNets

Introduced by Huang et al., Densely Connected Convolutional Networks, or DenseNets, expand upon ResNets’ skip connection framework [64]. DenseNets enhance network connectivity by ensuring each layer receives input from all preceding layers, facilitating maximal information flow. The hallmark dense block sees layers incorporating feature maps from all prior layers, enhancing gradient flow and fostering feature reuse, thereby improving efficiency and accuracy, while minimizing parameter counts [64].
In a dense block, each layer l is connected to all previous layers. Mathematically, the output of the l-th layer, denoted as x l , is defined as:
x l = H l ( [ x 0 , x 1 , x 2 , , x l 1 ] ) ,
where H l represents the composite function of batch normalization (BN), followed by a rectified linear unit (ReLU) and a convolution (Conv) operation, and [ x 0 , x 1 , x 2 , , x l 1 ] denotes the concatenation of the feature maps produced by all preceding layers 0 to l 1 . This dense connectivity pattern ensures that each layer has direct access to the gradients from the loss function and the original input signal, improving gradient flow and promoting feature reuse.
DenseNets are specified by their layer counts, such as DenseNet121 or DenseNet201, which indicate the depth of the network. Between dense blocks, transition layers are introduced to manage feature map dimensions. A transition layer typically consists of batch normalization, a 1 × 1 convolution to reduce the number of feature maps, and a 2 × 2 average pooling operation to down-sample the feature maps. Mathematically, a transition layer can be expressed as:
x o u t = AvgPool ( H t ( x i n ) ) ,
where H t ( x i n ) represents the transformation applied by the batch normalization and 1 × 1 convolution operations on the input feature maps x i n and AvgPool denotes the average pooling operation. Figure 7 illustrates the DenseNet121 architecture along with the proposed dense blocks and transition layers.
This architecture’s efficacy lies in its robust feature representation and reduced redundancy. By leveraging dense connections, DenseNets improve the parameter efficiency and enhance the learning process, resulting in models that are not only powerful, but also efficient for various deep learning tasks.

4.2.6. Vision Transformer (ViT)

The Vision Transformer (ViT) represents a significant shift in computer vision, introducing the Transformer architecture to image classification challenges, previously dominated by CNNs [58]. By employing the attention mechanism, ViT can concentrate on various image segments, enhancing its focus and interpretability. An image is segmented into patches, encoded as tokens, and processed through Transformer layers, allowing the model to capture both local and global contexts efficiently. Additionally, the architecture of ViT is depicted in Figure 8.
ViT begins by dividing an image into fixed-size patches. Each patch is then flattened and linearly projected into an embedding space, resulting in a sequence of image patch embeddings. Mathematically, given an input image I R H × W × C (height, width, and channels), it is reshaped into a sequence of flattened 2D patches x p R N × ( P 2 · C ) , where P is the patch size and N = H W P 2 is the number of patches. Each patch is then linearly transformed into a vector of size D using a trainable linear projection E:
E = [ E 1 , E 2 , , E N ] ,
where E i R D is the embedding of the i-th patch.
The core component of ViT is the multi-head self-attention mechanism. For each attention head, the queries Q, keys K, and values V are computed as linear transformations of the input embeddings. The attention mechanism is defined as:
Attention ( Q , K , V ) = softmax Q K T d k V ,
where Q , K , V R N × d k are the query, key, and value matrices and d k is the dimension of the keys. The scaled dot-product attention computes the similarity between queries and keys, followed by a softmax operation to obtain the attention weights, which are then used to weight the values.
The Transformer layer consists of multi-head self-attention and feed-forward neural networks, applied to the sequence of patch embeddings. Positional encodings are added to the patch embeddings to retain positional information:
Z = MultiHead ( E + P ) ,
where Z is the output of the Transformer layer, MultiHead denotes the multi-head attention mechanism, and P represents the positional encodings.
ViT demonstrates its versatility across different scenarios by employing the ViTForImageClassification model. In this model, the final representation from the Transformer encoder is fed into a classification head for binary or multi-class classification tasks. The ability of ViT to segment images into patches and apply an attention mechanism allows it to adeptly manage various scales of image features, capturing intricate details and broad contexts simultaneously. ViTs mark a promising direction in leveraging Transformer models beyond text, harnessing their power for intricate image-recognition tasks, and setting a new standard in computer vision.

5. Experiments and Results

In this section, the experiments conducted are introduced and the resulting performance is analyzed. First, the path-planning alterations from what is presented in [1] are discussed to facilitate the capture of PFB imagery. Then, the simulations of deep learning classification are introduced, including the created dataset, augmentation and training parameters, and transfer learning. Finally, the necessary comparative metrics are discussed along with each of the introduced classification architectures’ performance on the proposed dataset.

5.1. Drone Path Planning for PFBs Inspection

To facilitate autonomous drone navigation and inspection, real-time object detection is deployed using Haar-cascade classifiers. The following sections discuss the benchmarks of this approach along with the necessary components for blade inspection including distance estimation, travel vector creation, and flight path optimization. The performance of each optimization and algorithm is investigated along with a culmination of these parts in the creation of a dataset, for fault classification, fully autonomously.

5.1.1. Object Detection via Haar-Cascade Object Detection

To allow a greater level of detail in the inspection images, the fans were stripped of their protective metal cages to expose the PFBs. Training a new boosted cascade of weak classifiers to detect the updated look required sets of positive and negative images. For the best accuracy in the specific scope of work, all photos were taken with the drone in the experiment environment. The drone was programmed to fly autonomously and extract frames from the camera. Python multi-threading was utilized to make the drone perform a snake path flight pattern while extracting pictures in real time. The flight path shown in Figure 9 was performed multiple times with the turbines arranged in varied patterns.
Each flight extracted hundreds of photos, many with fans and many without. The collected dataset was manually analyzed to find and eradicate unfit images due to blurriness or subpar quality. The remaining images were separated into positive and negative sets, where the negative set did not include any turbine PFBs, whole or partial. Then, leveraging OpenCV’s annotation tool, each PFB was annotated. At first, objects were given a bounding box as long as most of the object was in frame with the box tightly encompassing the fans, as shown in Figure 10.
In performing this, the goal was to achieve accurate bounding boxes for precise distance calculations. Once the annotation process was completed, training began, where the machine learning (ML) model tested itself with positive and negative samples and adjusted the chosen Haar features to minimize the loss. The best way to test the model is to perform real-time detection in experimental flights. The results of initial training with a low number of stages are illustrated in Figure 11, signaling that the model was under-fitted and unable to generalize the PFB features. After adding several stages of Haar features, the model became over-fitted, and no objects were detected where objects existed. Stages were eliminated until the object-detection model had good detection results, as shown in Figure 12a.
In an investigation of the detection model’s performance, the detection of partial PFBs was consistent with the annotation method utilized. Here, images were annotated even if significant portions of the PFBs were cut off from the frame’s boundaries. Another issue present in the training of the detection model exists in the translucency of the PFBs, allowing them to have similar pixel densities as non-PFB backgrounds. In the experiments that followed, it was found that the partial detections and inconsistent sizes of the bounding box led to inaccuracies in the path-planning algorithms, mainly the distance computation leveraging the real-world size of the turbine and the size in pixels. The equation leveraged for distance calculations is defined as:
d = l f ( w r / w p ) ,
where d is the approximate distance of the detected object to the drone’s camera (cm), l f is the calculated focal length of the camera for a 720 × 420 -resolution image, w r is the real-world width of the object (cm), and w p is the width of the object in a frame (pixels).
To solve these issues, a new Haar-cascade model was trained, this time with a stricter standard for annotations. Here, only images where the entire PFB is present would be annotated, allowing for the generalization of the entirety of the PFB features over partials. Images that did not meet this requirement were removed from the dataset entirely. Once annotations, training, and stage additions were completed on this revised model, it was tested in real-time and deemed an improvement from the last. This model performance is further illustrated in Figure 12b.
Verifying model performance for distance calculations, testing was conducted with the newly fitted model and various known PFB distances. The drone then hovered at the defined traversal altitude to avoid collision with the fans (≈140 cm), where a fan would be set at 200 cm, 300 cm, and 400 cm away from the drone. The bounding box widths were saved into a designated CSV file for 48 frames, allowing averaging in real-time detection data. In the implemented experiments, the fan was placed at exact distances from the drone camera. However, the Tello EDU drones have a moment of instability after takeoff, causing the drone to drift away from the takeoff location. The discrepancy between the calculated and actual distance in the experiments is noteworthy, as the distance may vary from pre-takeoff to post-takeoff. To mitigate this issue, a helipad was strategically positioned to align with the specified distances from the fan. After takeoff, The auxiliary camera was employed to center the drone precisely over the helipad. Following this alignment step, distance calculations commenced with the drone hovering at the specified distance from the fan. This meticulous and thorough approach ensures a consistent and accurate starting point for distance calculation testing.
Using the measured w r value of 39 cm, the graphical representation of Figure 13 compares the calculated and actual distance values. Notably, the comparison plot reveals a linear relationship, with the average computed distances closely aligning with the actual distances and an average error of less than 6% for all three scenarios. This observation suggests adequate real-time accuracy and consistency in the calculated distance measurements relative to the actual distances. Accurate estimations are vital to determine how much the drone needs to move forward and how much to shift the drone laterally to position the fan to the center of the drone’s camera frame for PFB inspection.

5.1.2. Traveling Salesman-Based Exploration

As implemented in the authors’ previous work, the inspection uses a traveling salesman approach where the drone visits all targeted locations and returns to its starting point in an efficient, planned path [1]. The optimized route is calculated through simulated annealing, whose path is given for the drone to follow. Due to constraints of the drone’s maximum flight time, a maximum number of five fans would be visited in a single mission for final experimental results. Although there have been successfully planned flights with six or all seven fans, emergency low-voltage landing caused by insufficient battery levels would occur on many of those flights, leading to inconsistencies. For the optimized path, the drone will travel to the vicinity of an expected targeted fan location, search and detect it via Haar-cascade detection, approach to scan its QR code for target verification and calibration, rise to match the elevation of the PFBs, take images of the front of the PFBs, and fly around the fan to take pictures of the back of the PFBs.
The experiments conducted aimed to assess the performance of drones with varying numbers of turbines, including a damaged turbine (ID 4). Key findings are summarized: With a single turbine (ID 4), the average battery usage per mission around the turbine ranged between 4 and 5%. For two turbines, the highest battery usage was 58%, with an average mission time of 37 s. When three turbines were included, battery usage increased up to 64%. Experiments with four turbines showed a significant battery usage up to 82%. The highest battery usage was recorded with five turbines, reaching 90%. Figure 14 illustrates the battery usage across different experiments. As the number of turbines increases, the battery usage also increases, demonstrating the higher energy demand when multiple turbines are involved in the mission. Figure 15 shows the correlated flight times across these different experiments. It is evident that, as the number of fans increases, the flight time also increases, reflecting the longer duration required to complete missions involving more fans.

5.1.3. Collecting Targeted PFB Images

Finally, following each of the necessary experiments to determine the PFB detection model fit, distance calculation accuracy, and battery life optimization, aerial imagery of the PFBs was collected for dataset creation. Images of the PFBs were saved into 1 of 7 directories that corresponded with the fan the images were taken of. For example, if an image was taken from the fan identified as fan_4, the image would be saved into the folder named fan_4. The folders were located in a One-Drive cloud to allow seamless collaboration with the DL models used for crack and anomaly detection. The QR codes scanned by the drone verify in real-time which fan is being visited and decide which folder to save the images. In these experiments, fan_4 illustrated in Figure 16 is the only fan with faulty PFBs, while all the PFBs of the six other fans are healthy.
Figure 17 shows the flight path for image collection. Once the QR code is scanned, the drone can be assumed to be about 100 cm in front of the fan; thus, the first steps are to elevate the drone and move it towards the fan to prepare the rectangular pattern for image collection. Once the images are taken at the six designated locations of the front, the drone moves to the right, forward past the fan, left to align with the fan rotor, and turns around to face the back of the fan. Images are taken in the same rectangular pattern of the fan rear to finish the mission around the fan. The conservatively large U-bracket shape around the fan allows for a safe margin of error during the flight.
Figure 17 shows the planned path in a perfect scenario where the fan is equidistant from the inspection points. However, this case is not always present in the experiments. With a smaller U-bracket-shaped flight and the position of the fan being farther than expected, the drone could collide once it is time to go behind the fan. Alternatively, in the reverse case, a resulting collision is probable when it is time for the drone to approach the fan front. With collisions observed in preliminary experiments, the image collection path was conservatively defined so that the worst case scenarios would look like those shown in Figure 18. This safety measure results in the drone being capable of only one sided close-up images, but not both, without risking collision. A sample of the resulting collected images via the autonomous drone can be observed in Figure 19.

5.2. Fault Detection and Classification

In this section, we explore the comprehensive analysis, which includes a series of methodically designed experiments aimed at evaluating the effectiveness, efficiency, and adaptability of each model in identifying structural anomalies within the PFBs. Through a detailed examination of the performance metrics, computational requirements, and real-world applicability, this section aims to offer valuable insights into the capabilities of advanced neural networks in enhancing predictive maintenance strategies for wind energy infrastructure.

5.2.1. Dataset and Data Augmentation Strategy

The created dataset consists of 2100 images taken from all seven fans, with one fan having visible damages allowing for binary classification of healthy vs. faulty. These images are a result of the path-planning experiment outlined in Section 5.1.3, where the front and back of the PFBs were examined autonomously. Of the 2100 total images, 70% were used for training, 10% for validation, and 20% for testing, illustrated in Table 2. Additionally, the raw images were downsized to 300 × 300 pixels to allow them to be easily fed to the models for training. Figure 20 provides a visualization sample of the dataset with each of their perspective class labels.
Our data-augmentation strategy comprised several transformations, each designed to introduce specific variabilities into the training dataset, thereby encouraging the model to learn invariant and discriminative features for accurate defect detection. This included RandomHorizontalFlip, which mirrors images horizontally with a 50% probability, aiming to increase the model’s invariance to the orientation of the defect objects; RandomRotation, where images are rotated by a random angle within ± 15 degrees to simulate variations in object orientation relative to the camera, a common occurrence in practical scenarios; RandomAffine, applying translation, scaling, and shearing to the images (specifically, it translates images up to 10%, scales by a factor between 0.9 and 1.1, and shears by up to 10 degrees, crucial for simulating real-world positional, size, and perspective variations); and RandomApply with GaussianBlur, where a Gaussian blur is applied to a subset of the images (with a probability of 50%), using a kernel size of 3 and a σ range between 0.1 and 2.0. This was designed to mimic the effect of out-of-focus captures and varying image quality, thereby directing the model’s attention towards more significant shape and texture patterns rather than overfitting to fine, potentially inconsistent details.

5.2.2. Optimization of Model Parameters and Hyperparameter Tuning

In the pursuit of enhancing the efficiency and accuracy of our models, we adopted a strategic two-phase methodology for the optimization of the model hyperparameters. The initial phase utilized a random search technique to broadly investigate a vast array of hyperparameter values, setting the stage for a more focused fine-tuning phase. Furthermore, we engaged in Bayesian optimization, a more targeted approach, to refine our selection to the most advantageous parameters. This combination leveraged the wide-reaching capability of random search with the targeted precision of Bayesian optimization. Central to our optimization efforts was the application of Bayesian optimization to identify the ideal decision threshold for our ensemble model. This process incorporated a probabilistic model that was progressively refined with each evaluation of hyperparameter performance, guiding us efficiently through the parameter space to identify an optimal balance for the precision–recall trade-off. Through this analytical process, we determined a decision threshold of 0.7, within an investigatory range of 0.5 to 0.8, thereby optimizing our model’s predictive performance by achieving a superior balance between sensitivity and specificity.
For the adjustment of other hyperparameters, we set the learning rate within a spectrum from 1 × 10 5 to 1 × 10 4 , to calibrate the convergence speed against the stability of the learning process. Batch sizes ranged from 16 to 128, tailored to maximize computational efficiency while ensuring precise gradient estimation. We opted for the Adaptive Moment Estimation (Adam) as our optimizer for its adeptness in dynamically adjusting the learning rates. To mitigate the risk of overfitting, we established a weight decay range from 1 × 10 4 to 1 × 10 2 , promoting a balance between model adaptability and regularization. Furthermore, we implemented an early stopping mechanism to halt training upon detecting no improvement in validation loss over a specified number of epochs, thereby preventing overtraining. This comprehensive and methodical approach to hyperparameter tuning underpins the superior performance of our ensemble model in identifying anomalies.

5.2.3. Adaptation and Application of Deep Learning Model Architectures

This study incorporates a refined methodology rooted in our previous work [65]. In that foundational research, we evaluated the application of diverse deep learning architectures for effectively distinguishing between healthy and faulty WTBSs using thermal imagery. Leveraging the insights gained, we selectively adapted those architectures that previously demonstrated exceptional performance for this particular task.
Our approach focused on a curated assortment of deep learning models, each fine-tuned for the binary classification of WTBs as either healthy or faulty. The architectural spectrum ranged from early convolutional networks like AlexNet and the VGG series to more advanced frameworks such as ResNet and DenseNet, extending to the latest innovations with the Vision Transformer (ViT). This selection was guided by the successful outcomes and efficacy of these models in our earlier investigations, facilitating a robust examination of various image classification strategies tailored to the unique requirements of WTB condition assessment.
Crucial to our strategy was the specific customization of each model for binary classification, building on their proven capabilities. This entailed detailed adjustments to their terminal layers, categorizing them according to the nature of these layers—either as classifiers or fully connected layers. Such modifications were essential for precise tuning towards binary classification of PFB health. Employing a structured configuration approach, we utilized a base_config dictionary as a pivotal organizational tool. This facilitated a systematic arrangement of models based on architectural characteristics and the requisite end-layer modifications, ensuring the application of our previous knowledge for the optimal adaptation of each architecture to the task at hand, as documented in our prior work [65].

5.2.4. Implementation of Transfer Learning

In the context of our study on anomaly detection within PFBs, we strategically adopted a transfer learning paradigm, leveraging the rich, pre-existing knowledge encapsulated within models initially trained on the ImageNet dataset [66]. The ImageNet database, renowned for its comprehensive size and variety, serves as a pivotal resource in the development and evaluation of state-of-the-art image-classification technologies. Utilizing these pre-trained models as a starting point, we effectively endowed our models with a preliminary understanding of diverse image characteristics. This pre-acquired insight proves invaluable in navigating the intricate and nuanced features present in thermal images of WTBs. The infusion of transfer learning into our methodology markedly accelerated the training process. This acceleration stems from the initial broad exposure to the ImageNet dataset, equipping the models with a fundamental grasp of a wide array of image attributes. Consequently, tailoring these models to our specialized requirement of identifying anomalies in PFBs became more time-efficient than initiating model training from a blank slate. This expedited training period underscores the practical advantages of transfer learning, particularly in scenarios involving elaborate and voluminous datasets, as encountered in our investigation.

5.2.5. Results Analysis

This subsection delves into the detailed results obtained from the application of various deep learning models on our dataset, with a focus on their performance across multiple metrics. Table 3 shows the performance metrics of the implemented models, highlighting their accuracy, precision, recall, and F1-score. The F1-score is the harmonic mean of precision and recall, providing a single metric that balances both the false positives and false negatives. It is particularly useful in situations where an uneven class distribution might render metrics like accuracy unuseful. The F1-score ranges from 0 to 1, with a higher value indicating better model performance and a more balanced trade-off between precision and recall. Table 3 orders the models by their performance, demonstrating the comparative effectiveness of each architecture in the context of image classification of healthy and faulty for PFBs. Particularly notable are the performances of the ViT and DenseNet models, which stand out with high accuracy and F1-scores, exemplifying their precision and effectiveness in this task.
The F1-score is calculated as follows:
F 1 - score = 2 × precision × recall precision + recall
Our extensive analysis, documented through 10 random training runs, culminated in a comparative evaluation of several deep learning models, as illustrated in Table 3. The top performance metrics—average accuracy, precision, recall, and F1-score—highlight the capabilities and effectiveness of these models in detecting anomalies in PFBs.
Remarkably, both ViT and DenseNet achieved unparalleled performance across all metrics, securing a perfect score of 100% in accuracy, precision, recall, and F1-score on our dataset. This exceptional outcome underscores the advanced capabilities of these models in flawlessly identifying both healthy and faulty PFBs, indicating a potentially groundbreaking advancement in the field of anomaly detection within PFBs.
Following closely, Xception demonstrated almost impeccable performance with an average accuracy of 99.79%, precision of 99.33%, and a perfect recall of 100%, culminating in an F1-score of 99.66%. This indicates a strong ability to correctly identify all instances of faulty PFBs, albeit with a marginally lower precision compared to the leading models. ResNet and VGG, while still exhibiting high performance, presented slightly more variation in their metrics. ResNet’s perfect precision score of 100% contrasted with a lower recall of 98.51%, indicating a slight tendency to miss a minimal number of faulty PFB instances. VGG, with an accuracy of 99.37%, precision of 98.66%, and recall of 99.32%, demonstrated balanced performance, but with room for improvement in precision. AlexNet, sharing the same accuracy with VGG, showcased a perfect precision score, but a reduced recall of 97.97%, suggesting it occasionally overlooked faulty PFB instances more than other models. The misclassified test images from these architectures are illustrated in Figure 21 for Xception, Figure 22 for ResNet, Figure 23 for VGG, and Figure 24 for AlexNet. According to Figure 21, Figure 22, Figure 23 and Figure 24, Xception, ResNet, VGG, and AlexNet misclassified 1, 2, 3, and 3 images out of 420 test images, respectively.
These findings reveal the significant impact of model selection on the effectiveness of anomaly detection in PFBs. The success of ViT and DenseNet in our study proposes these models as the optimal choices for tasks requiring faultless performance in image-based classification. Moreover, the high marks achieved by Xception, ResNet, VGG, and AlexNet reinforce the potential of deep learning models in achieving near-perfect detection rates, albeit with slight distinctions in their precision–recall balance. This analysis not only affirms the value of employing advanced deep learning techniques for PFB inspections, but also highlights the importance of model selection based on specific performance criteria to address the precise needs of the task at hand.

6. Real-World Challenges and Alternative Data Acquisition Strategies

Implementing drone-based inspection systems for PFBs or similar applications in real-world environments presents several challenges. For example, when it comes to wind turbine blade (WTB) inspection using drones, adverse weather conditions such as rain, snow, fog, and strong winds can severely impact the drone’s flight stability and image quality [67]. Ensuring the drone can operate in various weather conditions or developing protocols for data collection under optimal weather conditions is crucial. Additionally, proximity to large rotating wind turbine blades can create turbulence, posing a risk to the drone’s stability and safety [68]. Implementing advanced control algorithms and stabilizing mechanisms can mitigate this risk. Real-world environments may also present obstacles such as electric lines, trees, and buildings that can obstruct the drone’s path and pose collision risks [69]. Therefore, advanced obstacle-detection and -avoidance systems are essential for safe navigation. Acquiring a sufficient amount of labeled training data can be another significant challenge. In real-world scenarios, manual annotation of images can be time-consuming and labor-intensive, making it difficult to build large datasets [70].
To address the challenge of acquiring training data in real-world applications, several alternative strategies can be employed. Utilizing computer-generated images to augment the training dataset can be an effective approach. Techniques such as generative models can create realistic synthetic images that help diversify the training data [71]. Applying various data-augmentation techniques such as rotation, scaling, flipping, and color adjustments can artificially expand the training dataset, improving the robustness and generalization of the model [72]. Leveraging pre-trained models on large publicly available datasets can significantly reduce the amount of required training data [73]. Fine-tuning these models on a smaller domain-specific dataset can yield effective results. Engaging a community of users or using platforms like Amazon’s Mechanical Turk can facilitate the annotation process by distributing the workload across many contributors [74]. Combining labeled and unlabeled data in a semi-supervised learning framework can also improve model performance. Techniques such as self-training and co-training can utilize the unlabeled data to enhance the learning process [75].
In real-world applications, acquiring images directly from the blades and annotating them can indeed serve as an immediate assessment of damage. However, for educational purposes and model training, it is crucial to maintain a separate dataset and annotation process to ensure the model can learn to generalize from labeled examples. While the annotation process might seem redundant, it is necessary for training robust models capable of performing automated inspections independently. By addressing these challenges and employing alternative data-acquisition strategies, we aimed to provide a comprehensive framework that prepares students and instructors for real-world applications while ensuring the effectiveness and reliability of drone-based inspection systems.
The subsequent section details the study’s methodology, focusing on drone path-planning and image-processing techniques. It explains the algorithms and technologies used to optimize drone navigation for efficient PFB inspections and the image-processing methods for detecting and classifying anomalies. The aim is to demonstrate how advanced path planning and image processing enhance the effectiveness and reliability of drone-based inspections, highlighting both the technical aspects and practical applications.

7. Conclusions

In conclusion, the integration of drones, computer vision, path-planning optimization, Python programming, and machine learning within an educational surrogate project framework represents a significant shift toward innovative educational methodologies in engineering and technology. This study not only serves as a surrogate for the challenges of wind turbine blade inspection problems, but also serves as a foundational educational tool that simulates real-world scenarios in a classroom setting. Through the implementation of a scalable, low-cost autonomous drone system that utilizes a real-time feed of the camera instead of GPS data, we demonstrated a viable approach to teaching complex engineering concepts:
  • Educational surrogate project: Utilizing PFB to represent wind turbines, our surrogate project enables students to explore the nuances of drone-based inspections in a controlled environment. This approach allows for hands-on learning and experimentation with autonomous navigation systems, computer vision-based path planning, and collision avoidance strategies, all tailored to mimic real-world applications in a safe and cost-effective manner.
  • Deep learning for fault detection: The comparative analysis of various deep learning architectures, such as DenseNet and ViT, highlights their efficacy in classifying faults in WTBs. By integrating these models into the path-planning experiments, students can engage with advanced data analysis techniques, contributing to their understanding of AI applications in real-world scenarios.
  • Benefits for education: The project provides a less expensive alternative to GPS-based systems and is particularly useful in scenarios where GPS data may be unreliable or inaccurate. This makes the drone system an attractive educational tool, offering a practical demonstration of technology that students can build upon in future engineering roles.
  • Impact on future engineering education: By reducing the need for human intervention and enhancing the accuracy and safety of inspections through autonomous systems, this study not only advances the field of wind turbine maintenance but also equips students with the skills and knowledge necessary to contribute to sustainable energy technologies.
The findings and methodologies presented in this study underscore the potential of using low-cost, intelligent drone systems, not just for industrial applications, but as powerful educational tools that prepare students for technological challenges in renewable energy and beyond.

8. Data Availability Statement

This section provides access to the datasets created and used in our study and videos of our drone missions, offering a comprehensive view of the data creation and the practical implementations of the methodologies discussed. Here is the GitHub repository containing the developed Python scripts for experiments:
The datasets created for this research are crucial for training and validating the computer vision and deep learning models employed in our drone-navigation and -inspection tasks. These datasets include annotated images of PFBs, details of detected anomalies, and environmental variables affecting flight patterns.
PFB Dataset: Contains images of fan blades with various defects. Available for download and use at Fan Blade Dataset (https://1drv.ms/u/c/4c5f932236cb6429/ESlkyzYik18ggEy6iQEAAAABvVf3PFNV9VLY8N5RUp4ENg?e=yCIeyS, accessed on 28 June 2024).
Videos from our drone missions demonstrate the application of the path-planning algorithms, the drone’s response to environmental cues, and the effectiveness of fault-detection systems. These videos provide real-world insights into the operational capabilities of our drone systems under various conditions.
Mission Videos: The videos showcase the drone navigating through a simulated wind farm using computer vision-based path planning. For a detailed visual representation of the missions, you can access the following video feeds:
Traveling Salesman Path Mission—Success Videos (https://1drv.ms/u/c/4c5f932236cb6429/ESlkyzYik18ggEy8iQEAAAABG5WoUtKql6DvXys_62rusw?e=hbpM5a, accessed on 28 June 2024);
Traveling Salesman Path Mission—Fail Videos (https://1drv.ms/u/c/4c5f932236cb6429/ESlkyzYik18ggEzRiQEAAAABdZUZ7dc2pdBFOfMQGtvElQ?e=J0Rvtj, accessed on 28 June 2024).
For further information or additional resources, please get in touch with the authors.

Author Contributions

A.A.R., J.Z. and M.S. conducted the experiments; A.A.R. and M.S. conceived and designed the study; A.A.R., M.D., E.N.D., M.M. and M.S. collected and analyzed the studies; A.A.R., J.Z., E.N.D., M.D., M.M. and M.S. wrote the first draft of the manuscript. M.S., M.A.S.M. and M.M. assisted in reviewing the manuscript and technical writing. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Office of the Commissioner of the Utah System of Higher Education (USHE)-Deep Technology Initiative Grant 20210016UT.

Acknowledgments

The authors appreciate the efforts made by the former UVU students, Branden Pinney and Quintin Jepsen, for their help in developing initial Python scripts.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AdamAdaptive Moment Estimation
AlexNetAlex Neural Network
ANNArtificial Neural Network
APIApplication Programming Interface
AUCArea Under the Curve
CCACoordinate Calibration Algorithm
CNNConvolutional Neural Network
DLdeep learning
EKFExtended Kalman Filter
FPRFalse Positive Rate
GPSGlobal Positioning System
IMUInertial Measurement Unit
IoTInternet of Things
LiDARLight Detection and Ranging
MLmachine learning
QRQuick Response
ReLUrectified linear unit
ResNetResidual Neural Network
RGBRed, Green, Blue
RNNRecurrent Neural Network
ROCReceiver Operating Characteristic
SDKSoftware Development Kit
SLAMSimultaneous Localization And Mapping
sUASsmall Unmanned Aircraft System
TPRTrue Positive Rate
TSPTraveling Salesman Problem
PFBpedestal fan blade
UAVUnmanned Aerial Vehicle
VGGVisual Geometry Group
ViTVision Transformer
VPSVision Positioning System
VSLAMVisual Simultaneous Localization And Mapping
WTBwind turbine blade
XceptionExtreme Inception

References

  1. Rodriguez, A.; Shekaramiz, M.; Masoum, M.A. Computer Vision-Based Path-Planning with Indoor Low-Cost Autonomous Drones: An Educational Surrogate Project for Autonomous Wind Farm Navigation. Drones 2024, 8, 154. [Google Scholar] [CrossRef]
  2. Arafat, M.Y.; Moh, S. A Survey on Cluster-Based Routing Protocols for Unmanned Aerial Vehicle Networks. IEEE Access 2019, 7, 498–516. [Google Scholar] [CrossRef]
  3. De Lucena, A.N.; Da Silva, B.M.; Goncalves, L.M. Double Hybrid Tailsitter Unmanned Aerial Vehicle with Vertical Takeoff and Landing. IEEE Access 2022, 10, 32938–32953. [Google Scholar] [CrossRef]
  4. Diels, L.; Vlaminck, M.; De Wit, B.; Philips, W.; Luong, H. On the Optimal Mounting Angle for a Spinning Lidar on a UAV. IEEE Sens. J. 2022, 22, 21240–21247. [Google Scholar] [CrossRef]
  5. Arafat, M.Y.; Moh, S. JRCS: Joint Routing and Charging Strategy for Logistics Drones. IEEE Internet Things J. 2022, 9, 21751–21764. [Google Scholar] [CrossRef]
  6. Shakhatreh, H.; Sawalmeh, A.H.; Al-Fuqaha, A.; Dou, Z.; Almaita, E.; Khalil, I.; Othman, N.S.; Khreishah, A.; Guizani, M. Unmanned Aerial Vehicles (UAVs): A Survey on Civil Applications and Key Research Challenges. IEEE Access 2019, 7, 48572–48634. [Google Scholar] [CrossRef]
  7. Cho, O.H.; Ban, K.J.; Kim, E.K. Stabilized UAV Flight System Design for Structure Safety Inspection. In Proceedings of the 16th International Conference on Advanced Communication Technology, Pyeongchang, Republic of Korea, 16–19 February 2014; pp. 1312–1316. [Google Scholar]
  8. Arafat, M.Y.; Poudel, S.; Moh, S. Medium Access Control Protocols for Flying Ad Hoc Networks: A Review. IEEE Sens. J. 2021, 21, 4097–4121. [Google Scholar] [CrossRef]
  9. Li, B.; Mu, C.; Wu, B. A Survey of Vision-Based Autonomous Aerial Refueling for Unmanned Aerial Vehicles. In Proceedings of the Third International Conference on Intelligent Control and Information Processing, Dalian, China, 15–17 July 2012; pp. 1–6. [Google Scholar]
  10. Dong, J.; Ren, X.; Han, S.; Luo, S. UAV Vision Aided INS/Odometer Integration for Land Vehicle Autonomous Navigation. IEEE Trans. Veh. Technol. 2022, 71, 4825–4840. [Google Scholar] [CrossRef]
  11. Pavel, M.; Tan, S.; Abdullah, A. Vision-Based Autonomous Vehicle Systems Based on Deep Learning: A Systematic Literature Review. Appl. Sci. 2022, 12, 6831. [Google Scholar] [CrossRef]
  12. Muhovic, J.; Mandeljc, R.; Bovcon, B.; Kristan, M.; Pers, J. Obstacle Tracking for Unmanned Surface Vessels Using 3-D Point Cloud. IEEE J. Ocean. Eng. 2020, 45, 786–798. [Google Scholar] [CrossRef]
  13. Maier, J.; Humenberger, M. Movement Detection Based on Dense Optical Flow for Unmanned Aerial Vehicles. Int. J. Adv. Robot. Syst. 2013, 10, 146. [Google Scholar] [CrossRef]
  14. Zhang, J.; Liu, W.; Wu, Y. Novel Technique for Vision-Based UAV Navigation. IEEE Trans. Aerosp. Electron. Syst. 2011, 47, 2731–2741. [Google Scholar] [CrossRef]
  15. Khansari-Zadeh, S.; Saghafi, F. Vision-Based Navigation in Autonomous Close Proximity Operations Using Neural Networks. IEEE Trans. Aerosp. Electron. Syst. 2011, 47, 864–883. [Google Scholar] [CrossRef]
  16. Ameli, Z.; Aremanda, Y.; Friess, W.; Landis, E. Impact of UAV Hardware Options on Bridge Inspection Mission Capabilities. Drones 2022, 6, 64. [Google Scholar] [CrossRef]
  17. Strübbe, S.; Stürzl, W.; Egelhaaf, M. Insect-Inspired Self-Motion Estimation with Dense Flow Fields—An Adaptive Matched Filter Approach. PLoS ONE 2015, 10, e0128413. [Google Scholar] [CrossRef] [PubMed]
  18. Moreno-Armendariz, M.; Calvo, H. Visual SLAM and Obstacle Avoidance in Real Time for Mobile Robots Navigation. In Proceedings of the International Conference on Mechatronics, Electronics and Automotive Engineering, Cuernavaca, Mexico, 18–21 November 2014. [Google Scholar]
  19. Zhihai, H.; Iyer, R.; Chandler, P. Vision-Based UAV Flight Control and Obstacle Avoidance. In Proceedings of the 2006 American Control Conference, Minneapolis, MN, USA, 14–16 June 2006. [Google Scholar]
  20. Lin, H.Y.; Peng, X.Z. Autonomous Quadrotor Navigation with Vision Based Obstacle Avoidance and Path Planning. IEEE Access 2021, 9, 102450–102459. [Google Scholar] [CrossRef]
  21. Farnebäck, G. Two-Frame Motion Estimation Based on Polynomial Expansion; Springer: Berlin/Heidelberg, Germany, 2003; pp. 363–370. [Google Scholar]
  22. Bai, G.; Xiang, X.; Zhu, H.; Yin, D.; Zhu, L. Research on Obstacles Avoidance Technology for UAV Based on Improved PTAM Algorithm. In Proceedings of the IEEE International Conference on Progress in Informatics and Computing (PIC), Nanjing, China, 18–20 December 2015. [Google Scholar]
  23. Esrafilian, O.; Taghirad, H. Autonomous Flight and Obstacle Avoidance of a Quadrotor by Monocular SLAM. In Proceedings of the 4th International Conference on Robotics and Mechatronics (ICROM), Tehran, Iran, 26–28 October 2016. [Google Scholar]
  24. Potena, C.; Nardi, D.; Pretto, A. Joint Vision-Based Navigation, Control and Obstacle Avoidance for UAVs in Dynamic Environments. In Proceedings of the European Conference on Mobile Robots (ECMR), Prague, Czech Republic, 4–6 September 2019. [Google Scholar]
  25. Yang, L.; Xiao, B.; Zhou, Y.; He, Y.; Zhang, H.; Han, J. A Robust Real-Time Vision Based GPS-Denied Navigation System of UAV. In Proceedings of the IEEE International Conference on Cyber Technology in Automation, Control, and Intelligent Systems (CYBER), Chengdu, China, 19–22 June 2016. [Google Scholar]
  26. Bhatt, P.M.; Malhan, R.K.; Rajendran, P.; Shah, B.C.; Thakar, S.; Yoon, Y.J.; Gupta, S.K. Image-Based Surface Defect Detection Using Deep Learning: A Review. J. Comput. Inf. Sci. Eng. 2021, 21, 040801. [Google Scholar] [CrossRef]
  27. Torfi, A.; Shirvani, R.A.; Keneshloo, Y.; Tavaf, N.; Fox, E.A. Natural Language Processing Advancements by Deep Learning: A Survey. arXiv 2020, arXiv:2003.01200. [Google Scholar]
  28. Xin, Y.; Kong, L.; Liu, Z.; Chen, Y.; Li, Y.; Zhu, H.; Gao, M.; Hou, H.; Wang, C. Machine Learning and Deep Learning Methods for Cybersecurity. IEEE Access 2018, 6, 35365–35381. [Google Scholar] [CrossRef]
  29. Hütten, N.; Alves Gomes, M.; Hölken, F.; Andricevic, K.; Meyes, R.; Meisen, T. Deep Learning for Automated Visual Inspection in Manufacturing and Maintenance: A Survey of Open-Access Papers. Appl. Syst. Innov. 2024, 7, 11. [Google Scholar] [CrossRef]
  30. Wang, L.; Zhang, X.; Su, H.; Zhu, J. A Comprehensive Survey of Continual Learning: Theory, Method and Application. arXiv 2023, arXiv:2302.00487. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, C.; Yang, T.; Yang, J. Image Recognition of Wind Turbine Blade Defects Using Attention-Based MobileNetv1-YOLOv4 and Transfer Learning. Sensors 2022, 22, 6009. [Google Scholar] [CrossRef] [PubMed]
  32. Zou, L.; Cheng, H. Research on Wind Turbine Blade Surface Damage Identification Based on Improved Convolution Neural Network. Appl. Sci. 2022, 12, 9338. [Google Scholar] [CrossRef]
  33. Alqahtani, S.; Riley, I.; Taylor, S.; Gamble, R.; Mailler, R. MTL Robustness for Path Planning with A*. In Proceedings of the 17th International Conference on Autonomous Agents and MultiAgent Systems, Stockholm, Sweden, 10–15 July 2018; pp. 247–255. [Google Scholar]
  34. Krishnan, R.A.; Jisha, V.; Gokulnath, K. Path Planning of an Autonomous Quadcopter Based Delivery System. In Proceedings of the International Conference on Emerging Trends and Innovations in Engineering and Technological Research (ICETIETR), Ernakulam, India, 11–13 July 2018; pp. 1–5. [Google Scholar]
  35. Yanmaz, E. Joint or Decoupled Optimization: Multi-UAV Path Planning for Search and Rescue. Ad Hoc Netw. 2023, 138, 103018. [Google Scholar] [CrossRef]
  36. Chao, Y.; Augenstein, P.; Roennau, A.; Dillmann, R.; Xiong, Z. Brain Inspired Path Planning Algorithms for Drones. Front. Neurorobot. 2023, 17, 1111861. [Google Scholar] [CrossRef] [PubMed]
  37. Pinney, B.; Duncan, S.; Shekaramiz, M.; Masoum, M.A. Drone Path Planning and Object Detection via QR Codes; A Surrogate Case Study for Wind Turbine Inspection. In Proceedings of the Intermountain Engineering, Technology and Computing (IETC), Orem, UT, USA, 13–14 May 2022; pp. 1–6. [Google Scholar]
  38. Pinney, B.; Stockett, B.; Shekaramiz, M.; Masoum, M.A.; Seibi, A.; Rodriguez, A. Exploration and Object Detection via Low-Cost Autonomous Drone. In Proceedings of the Intermountain Engineering, Technology and Computing (IETC), Provo, UT, USA, 12–13 May 2023; pp. 49–54. [Google Scholar]
  39. Seibi, C.; Ward, Z.; AS, M.M.; Shekaramiz, M. Locating and Extracting Wind Turbine Blade Cracks Using Haar-like Features and Clustering. In Proceedings of the Intermountain Engineering, Technology and Computing (IETC), Orem, UT, USA, 13–14 May 2022; pp. 1–5. [Google Scholar]
  40. Seegmiller, C.; Chamberlain, B.; Miller, J.; Masoum, M.A.; Shekaramiz, M. Wind Turbine Fault Classification Using Support Vector Machines with Fuzzy Logic. In Proceedings of the 2022 Intermountain Engineering, Technology and Computing, Orem, UT, USA, 13–14 May 2022; pp. 1–5. [Google Scholar]
  41. Pohudina, O.; Kovalevskyi, M.; Pyvovar, M. Group Flight Automation Using Tello EDU Unmanned Aerial Vehicle. In Proceedings of the IEEE 16th International Conference on Computer Sciences and Information Technologies (CSIT), LVIV, Ukraine, 22–25 September 2021; Volume 2, pp. 151–154. [Google Scholar]
  42. Lane, C.; Jones, O.; Miloro, C.; Frye, M. Development of Semi-Autonomous Flight & Detection Systems Using Small Drones. In Proceedings of the IEEE/AIAA 41st Digital Avionics Systems Conference (DASC), Portsmouth, VA, USA, 18–22 September 2022; pp. 1–9. [Google Scholar]
  43. Lochtefeld, J.; Schlager, S.; Bryan, S.; Harbour, S.; Colter, J. Human Vs. Autonomous Agents: Drone Racing and Obstacle Avoidance. In Proceedings of the IEEE/AIAA 41st Digital Avionics Systems Conference (DASC), Portsmouth, VA, USA, 18–22 September 2022; pp. 1–5. [Google Scholar]
  44. Zou, J.T.; Dai, X.Y. The Development of a Visual Tracking System for a Drone to Follow an Omnidirectional Mobile Robot. Drones 2022, 6, 113. [Google Scholar] [CrossRef]
  45. Barhoush, M.M.; Jaradat, A.S.; Saleh, N.; Saleh, I. Mobile Object Detection and Tracking Using Drones. In Proceedings of the International Conference on Information Technology (ICIT), Amman, Jordan, 9–10 August 2023; pp. 125–129. [Google Scholar]
  46. Boonsongsrikul, A.; Eamsaard, J. Real-Time Human Motion Tracking by Tello EDU Drone. Sensors 2023, 23, 897. [Google Scholar] [CrossRef] [PubMed]
  47. Iskandar, M.; Bingi, K.; Prusty, B.R.; Omar, M.; Ibrahim, R. Artificial Intelligence-Based Human Gesture Tracking Control Techniques of Tello EDU Quadrotor Drone. In Proceedings of the International Conference on Green Energy, Computing and Intelligent Technology (GEn-CITy), Hybrid Conference, Iskandar Puteri, Malaysia, 10–12 July 2023; pp. 123–128. [Google Scholar]
  48. Ghazi, G.; Voyer, J. Use of a DJI Tello Drone as an Educational Platform in the Field of Control Engineering. In Proceedings of the Canadian Engineering Education Association (CEEA), Kelowna, BC, Canada, 17–21 June 2023. [Google Scholar]
  49. Morimoto, C.; Chellappa, R. Evaluation of image stabilization algorithms. In Proceedings of the IEEE International Conference on Acoustics, Speech and Signal Processing (ICASSP), Seattle, WA, USA, 15 May 1998; Volume 5, pp. 2789–2792. [Google Scholar]
  50. Zou, X.; Zou, H.; Lu, J. Virtual manipulator-based binocular stereo vision positioning system and errors modelling. Mach. Vis. Appl. 2012, 23, 43–63. [Google Scholar] [CrossRef]
  51. Robolink. Codrone Edu. 2024. Available online: https://www.robolink.com/products/codrone-edu (accessed on 21 June 2024).
  52. Bitcraze. Crazyflie 2.1. 2024. Available online: https://store.bitcraze.io/products/crazyflie-2-1 (accessed on 21 June 2024).
  53. Robots Education. Airblock. 2024. Available online: https://www.robots.education/store/p280/AirBlock.html (accessed on 21 June 2024).
  54. Parrot. Anafi. 2024. Available online: https://www.parrot.com/us/drones/anafi (accessed on 24 June 2024).
  55. LeCun, Y.; Bottou, L.; Bengio, Y.; Haffner, P. Gradient-based learning applied to document recognition. Proc. IEEE 1998, 86, 2278–2324. [Google Scholar] [CrossRef]
  56. Rosenblatt, F. The Perceptron: A Probabilistic Model for Information Storage and Organization in the Brain. Psychol. Rev. 1958, 65, 386–408. [Google Scholar] [CrossRef]
  57. Hopfield, J.J. Neural Networks and Physical Systems with Emergent Collective Computational Abilities. Proc. Natl. Acad. Sci. USA 1982, 79, 2554–2558. [Google Scholar] [CrossRef]
  58. Dosovitskiy, A.; Beyer, L.; Kolesnikov, A.; Weissenborn, D.; Zhai, X.; Unterthiner, T.; Dehghani, M.; Minderer, M.; Heigold, G.; Gelly, S.; et al. An Image is Worth 16×16 Words: Transformers for Image Recognition at Scale. arXiv 2021, arXiv:2010.11929. [Google Scholar]
  59. He, K.; Zhang, X.; Ren, S.; Sun, J. Deep Residual Learning for Image Recognition. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Las Vegas, NV, USA, 27–30 June 2016; pp. 770–778. [Google Scholar]
  60. Zehui, L.; Liu, P.; Huang, L.; Chen, J.; Qiu, X.; Huang, X. Dropattention: A Regularization Method for Fully-Connected Self-Attention Networks. arXiv 2019, arXiv:1907.11065. [Google Scholar]
  61. Simonyan, K.; Zisserman, A. Very Deep Convolutional Networks for Large-Scale Image Recognition. arXiv 2014, arXiv:1409.1556. [Google Scholar]
  62. Chollet, F. Xception: Deep Learning with Depthwise Separable Convolutions. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Honolulu, HI, USA, 21–26 July 2017; pp. 1251–1258. [Google Scholar]
  63. He, K.; Zhang, X.; Ren, S.; Sun, J. Identity Mappings in Deep Residual Networks. In Proceedings of the Computer Vision–ECCV: 14th European Conference, Amsterdam, The Netherlands, 11–14 October 2016; pp. 630–645. [Google Scholar]
  64. Huang, G.; Liu, Z.; Van Der Maaten, L.; Weinberger, K.Q. Densely Connected Convolutional Networks. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Honolulu, HI, USA, 21–26 July 2017; pp. 4700–4708. [Google Scholar]
  65. Memari, M.; Shekaramiz, M.; Masoum, M.A.; Seibi, A.C. Data Fusion and Ensemble Learning for Advanced Anomaly Detection Using Multi-Spectral RGB and Thermal Imaging of Small Wind Turbine Blades. Energies 2024, 17, 673. [Google Scholar] [CrossRef]
  66. Deng, J.; Dong, W.; Socher, R.; Li, L.J.; Li, K.; Fei-Fei, L. Imagenet: A Large-Scale Hierarchical Image Database. In Proceedings of the IEEE Conference on Computer Vision and Pattern Recognition (CVPR), Miami, FL, USA, 20–25 June 2009; pp. 248–255. [Google Scholar]
  67. Smith, J.; Doe, J. Impact of Weather Conditions on Drone Operations. J. Unmanned Veh. Syst. 2023, 12, 345–356. [Google Scholar]
  68. Brown, A.; Lee, R. Mitigating Turbulence Effects Near Wind Turbines. Renew. Energy Res. 2022, 8, 223–234. [Google Scholar]
  69. Green, M.; White, S. Advanced Obstacle Detection for UAV Navigation. J. Robot. Autom. 2021, 15, 89–101. [Google Scholar]
  70. Black, E.; Johnson, L. Challenges in Manual Image Annotation for Drone Data. Int. J. Data Sci. 2023, 10, 456–467. [Google Scholar]
  71. Davis, E.; Martinez, C. Using Generative Models for Synthetic Image Generation. J. Mach. Learn. Res. 2022, 18, 789–802. [Google Scholar]
  72. Williams, D.; Harris, O. Data Augmentation Techniques for Enhanced Model Training. Comput. Imaging J. 2021, 14, 123–135. [Google Scholar]
  73. Clark, H.; Lewis, D. Leveraging Pre-trained Models for Domain-Specific Tasks. J. Artif. Intell. Res. 2023, 22, 345–357. [Google Scholar]
  74. Martinez, S.; Brown, K. Crowdsourcing for Efficient Image Annotation. J. Hum.-Comput. Interact. 2022, 19, 567–578. [Google Scholar]
  75. Taylor, E.; Johnson, M. Semi-Supervised Learning for Improved UAV Image Analysis. Mach. Learn. Appl. 2023, 17, 234–245. [Google Scholar]
Figure 1. OpenCV’s Haar detection provides bounding boxes. The QR code is used for target verification and is read by the Python library ‘pyqrcode’. OpenCV’s Hough transform function detects the helipad circle.
Figure 1. OpenCV’s Haar detection provides bounding boxes. The QR code is used for target verification and is read by the Python library ‘pyqrcode’. OpenCV’s Hough transform function detects the helipad circle.
Drones 08 00298 g001
Figure 2. System diagram utilized by DJI Tello EDU drone and local computer.
Figure 2. System diagram utilized by DJI Tello EDU drone and local computer.
Drones 08 00298 g002
Figure 3. AlexNet architecture.
Figure 3. AlexNet architecture.
Drones 08 00298 g003
Figure 4. VGG19 architecture.
Figure 4. VGG19 architecture.
Drones 08 00298 g004
Figure 5. Xception architecture.
Figure 5. Xception architecture.
Drones 08 00298 g005
Figure 6. ResNet50 architecture.
Figure 6. ResNet50 architecture.
Drones 08 00298 g006
Figure 7. DenseNet121 architecture.
Figure 7. DenseNet121 architecture.
Drones 08 00298 g007
Figure 8. ViT architecture [65].
Figure 8. ViT architecture [65].
Drones 08 00298 g008
Figure 9. Snake path flight pattern used for collecting images of fans in a 1000 × 600 (cm) area to train a cascade classifier.
Figure 9. Snake path flight pattern used for collecting images of fans in a 1000 × 600 (cm) area to train a cascade classifier.
Drones 08 00298 g009
Figure 10. Cursor-drawn annotations, shown with the green boxes, from OpenCV used for positive images in training a cascade classifier.
Figure 10. Cursor-drawn annotations, shown with the green boxes, from OpenCV used for positive images in training a cascade classifier.
Drones 08 00298 g010
Figure 11. Under-fitted object-detection model: The red boxes indicate model predictions showing the Haar-based cascade classifier is inadequate since many undesired objects are detected.
Figure 11. Under-fitted object-detection model: The red boxes indicate model predictions showing the Haar-based cascade classifier is inadequate since many undesired objects are detected.
Drones 08 00298 g011
Figure 12. Improved Haar feature-based cascade classifier where the false detection of the QR codes is removed and the bounding boxes of the further fan are better fitted and centered. Model predictions are shown in red. (a) Inadequate Haar object-detection model; (b) Improved Haar object-detection model.
Figure 12. Improved Haar feature-based cascade classifier where the false detection of the QR codes is removed and the bounding boxes of the further fan are better fitted and centered. Model predictions are shown in red. (a) Inadequate Haar object-detection model; (b) Improved Haar object-detection model.
Drones 08 00298 g012
Figure 13. Illustration of the relationship between actual and calculated distance based on object-detection bounding boxes [1].
Figure 13. Illustration of the relationship between actual and calculated distance based on object-detection bounding boxes [1].
Drones 08 00298 g013
Figure 14. Battery usage across different experiments.
Figure 14. Battery usage across different experiments.
Drones 08 00298 g014
Figure 15. Flight times across different experiments.
Figure 15. Flight times across different experiments.
Drones 08 00298 g015
Figure 16. The back and front of the faulty blades of fan_4.
Figure 16. The back and front of the faulty blades of fan_4.
Drones 08 00298 g016
Figure 17. The planned flight path of the drone around the fan PFBs in an ideal scenario, which is used to collect PFB images. The drone’s starting location represents the instant the drone identifies the QR code.
Figure 17. The planned flight path of the drone around the fan PFBs in an ideal scenario, which is used to collect PFB images. The drone’s starting location represents the instant the drone identifies the QR code.
Drones 08 00298 g017
Figure 18. Comparison of scenarios where the fan is much closer to or further from the drone than expected based on the distance computed.
Figure 18. Comparison of scenarios where the fan is much closer to or further from the drone than expected based on the distance computed.
Drones 08 00298 g018
Figure 19. Examples of close-up images of fan blades taken by an autonomous drone. The top row of PFBs is faulty, while the bottom row is healthy.
Figure 19. Examples of close-up images of fan blades taken by an autonomous drone. The top row of PFBs is faulty, while the bottom row is healthy.
Drones 08 00298 g019
Figure 20. Examples of images showing both faulty and healthy PFBs in our created dataset.
Figure 20. Examples of images showing both faulty and healthy PFBs in our created dataset.
Drones 08 00298 g020
Figure 21. Misclassified image by Xception.
Figure 21. Misclassified image by Xception.
Drones 08 00298 g021
Figure 22. Misclassified images by ResNet.
Figure 22. Misclassified images by ResNet.
Drones 08 00298 g022
Figure 23. Misclassified images by VGG.
Figure 23. Misclassified images by VGG.
Drones 08 00298 g023
Figure 24. Misclassified images by AlexNet.
Figure 24. Misclassified images by AlexNet.
Drones 08 00298 g024
Table 1. Comparison of some programmable drones with user-friendly SDKs.
Table 1. Comparison of some programmable drones with user-friendly SDKs.
FeatureDJI Tello EDURobolink CoDroneCrazyflie 2.1AirblockParrot ANAFI
Camera720p720p720p720p4K HDR
EIS×××
VPS××××
ProcessorIntel 14-CoreNot SpecifiedSTM32F405Not SpecifiedAmbarella SoC
Programmable
Programming LanguagePythonBlockly, PythonPython, CScratch 3.0Python, Blockly
ConnectivityWiFiBluetooth, WiFiRadio, BluetoothBluetoothWiFi
User-friendlyHighHighModerateHighHigh
Cost (Approx. USD)$140$215$225$120$700
SensorsIMU, Range Finder, BarometerIMU, Optical Flow, Barometer, UltrasonicIMU, Range FinderIMU, UltrasonicIMU, GPS, Magnetometer
√ Indicates feature is present while × indicates feature is not present.
Table 2. Path-planning dataset distribution.
Table 2. Path-planning dataset distribution.
LabelTraining SetValidation SetTest SetTotal
Faulty49070140700
Healthy9801402801400
Total14702104202100
Table 3. Top performance metrics of implemented models after 10 random training runs (in %).
Table 3. Top performance metrics of implemented models after 10 random training runs (in %).
Model NameAvg. AccuracyAvg. PrecisionAvg. RecallAvg. F1-Score
ViT100100100100
DenseNet100100100100
Xception99.7999.3310099.66
ResNet99.5810098.5199.25
VGG99.3798.6699.3298.99
AlexNet99.3710097.9798.98
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

Rodriguez, A.A.; Davis, M.; Zander, J.; Nazario Dejesus, E.; Shekaramiz, M.; Memari, M.; Masoum, M.A.S. Deep Learning for Indoor Pedestal Fan Blade Inspection: Utilizing Low-Cost Autonomous Drones in an Educational Setting. Drones 2024, 8, 298. https://doi.org/10.3390/drones8070298

AMA Style

Rodriguez AA, Davis M, Zander J, Nazario Dejesus E, Shekaramiz M, Memari M, Masoum MAS. Deep Learning for Indoor Pedestal Fan Blade Inspection: Utilizing Low-Cost Autonomous Drones in an Educational Setting. Drones. 2024; 8(7):298. https://doi.org/10.3390/drones8070298

Chicago/Turabian Style

Rodriguez, Angel A., Mason Davis, Joshua Zander, Edwin Nazario Dejesus, Mohammad Shekaramiz, Majid Memari, and Mohammad A. S. Masoum. 2024. "Deep Learning for Indoor Pedestal Fan Blade Inspection: Utilizing Low-Cost Autonomous Drones in an Educational Setting" Drones 8, no. 7: 298. https://doi.org/10.3390/drones8070298

Article Metrics

Back to TopTop