[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (157)

Search Parameters:
Keywords = blade crack

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 10666 KiB  
Article
Prediction of Mechanical Properties and Fracture Behavior of TC17 Linear Friction Welded Joint Based on Finite Element Simulation
by Xuan Xiao, Yue Mao and Li Fu
Materials 2025, 18(1), 128; https://doi.org/10.3390/ma18010128 - 31 Dec 2024
Viewed by 328
Abstract
TC17 titanium alloy is widely used in the aviation industry for dual-performance blades, and linear friction welding (LFW) is a key technology for its manufacturing and repair. However, accurate evaluation of the mechanical properties of TC17−LFW joints and research on their joint fracture [...] Read more.
TC17 titanium alloy is widely used in the aviation industry for dual-performance blades, and linear friction welding (LFW) is a key technology for its manufacturing and repair. However, accurate evaluation of the mechanical properties of TC17−LFW joints and research on their joint fracture behavior are still not clear. Therefore, this paper used the finite element numerical simulation method (FEM) to investigate the mechanical behavior of the TC17−LFW joint with a complex micro−structure during the tensile processing, and predicted its mechanical properties and fracture behavior. The results indicate that the simulated elastic modulus of the joint is 108.5 GPa, the yield strength is 1023.2 MPa, the tensile strength is 1067.5 MPa, and the elongation is 1.98%. The deviations from measured results between simulated results are less than 2%. The stress and strain field studies during the processing show that the material located at the upper and lower edges of the joint in the WZ experiences stress and strain concentration, followed by the extending of the stress and strain concentration zone toward the center of the WZ. And finally, the strain concentration zone covered the entire WZ. The fracture behavior studies show that the material necking occurs in the TMAZ of TC17(α + β) and WZ, while cracks first appear in the WZ. Subsequently, joint cracks propagate along the TC17(α + β) side of the WZ until fracture occurs. There are obvious tearing edges formed by the partial tearing of the WZ structure in the simulated fracture surface, and there are fracture surfaces with different height differences at the center of the joint crack, indicating that the joint has mixed fracture characteristics. Full article
(This article belongs to the Special Issue Advanced Materials Joining and Manufacturing Techniques)
Show Figures

Figure 1

Figure 1
<p>Diagram flow of the methodology and investigation steps.</p>
Full article ">Figure 2
<p>LFW machine setup and joint sample: (<b>a</b>) machine; (<b>b</b>) operation mode; (<b>c</b>) TC17−LFW joint.</p>
Full article ">Figure 3
<p>Morphology of TC17−LFW Joint.</p>
Full article ">Figure 4
<p>Schematic diagram of tensile specimen.</p>
Full article ">Figure 5
<p>Geometric model and mesh of tensile processing in TC17−LFW joint: (<b>a</b>) geometric model; (<b>b</b>) mesh.</p>
Full article ">Figure 6
<p>Stress–strain curves of different areas of TC17−LFW joint obtained by inverse material parameters using nanoindentation load displacement curves.</p>
Full article ">Figure 7
<p>Material parameter settings of tensile processing model in TC17−LFW joint.</p>
Full article ">Figure 8
<p>Boundary conditions of tensile processing model in TC17−LFW joint.</p>
Full article ">Figure 9
<p>Boundary conditions of tensile processing model in TC17−LFW joint.</p>
Full article ">Figure 10
<p>Stress field during tensile processing of TC17−LFW joint: (<b>a</b>) after yielding; (<b>b</b>) deformed 45%; (<b>c</b>) deformed 60%; (<b>d</b>) Deformed 90%.</p>
Full article ">Figure 11
<p>Strain field during tensile processing of TC17−LFW joint: (<b>a</b>) after yielding; (<b>b</b>) deformed 45%; (<b>c</b>) deformed 60%; (<b>d</b>) deformed 90%.</p>
Full article ">Figure 12
<p>Stress field during tensile processing of surface XOY of TC17−LFW joint: (<b>a</b>) deformed 45%; (<b>b</b>) deformed 60%; (<b>c</b>) deformed 90%.</p>
Full article ">Figure 13
<p>Strain field during tensile processing of surface XOY of TC17−LFW joint: (<b>a</b>) deformed 45%; (<b>b</b>) deformed 60%; (<b>c</b>) deformed 90%.</p>
Full article ">Figure 14
<p>DIC full-field strain during tensile processing of TC17−LFW joint: (<b>a</b>) deformed 45%; (<b>b</b>) deformed 60%; (<b>c</b>) deformed 90%.</p>
Full article ">Figure 15
<p>Neck shrinkage deformation and fracture behavior during tensile processing of TC17−LFW joint: (<b>a</b>) initiation; (<b>b</b>) neck shrinkage; (<b>c</b>) cracking; (<b>d</b>) fracture.</p>
Full article ">Figure 16
<p>Simulation fracture morphology of TC17−LFW joint.</p>
Full article ">Figure 17
<p>Fracture morphology of TC17−LFW joint: (<b>a</b>) overall morphology; (<b>b</b>–<b>d</b>) local fracture morphology.</p>
Full article ">
19 pages, 11286 KiB  
Article
Lifetime Prediction of Single Crystal Nickel-Based Superalloys
by Çağatay Kasar, Utku Kaftancioglu, Emin Bayraktar and Ozgur Aslan
Appl. Sci. 2025, 15(1), 201; https://doi.org/10.3390/app15010201 - 29 Dec 2024
Viewed by 353
Abstract
Single crystal nickel-based superalloys are extensively used in turbine blade applications due to their superior creep resistance compared to their polycrystalline counterparts. With the high creep resistance, high cycle fatigue (HCF) and low cycle fatigue (LCF) become primary failure mechanisms for such applications. [...] Read more.
Single crystal nickel-based superalloys are extensively used in turbine blade applications due to their superior creep resistance compared to their polycrystalline counterparts. With the high creep resistance, high cycle fatigue (HCF) and low cycle fatigue (LCF) become primary failure mechanisms for such applications. This study investigates the fatigue life prediction of CMSX-4 using a combination of crystal plasticity and lifetime assessment models. The constitutive crystal plasticity model simulates the anisotropic, rate-dependent deformation behavior of CMSX-4, while the modified Chaboche damage model is used for lifetime assessment, focusing on cleavage stresses on active slip planes to include anisotropy. Both qualitative and quantitative data obtained from HCF experiments on single crystal superalloys with notched geometry were used for validation of the model. Furthermore, artificial neural networks (ANNs) were employed to enhance the accuracy of lifetime predictions across varying temperatures by analyzing the fatigue curves obtained from the damage model. The integration of crystal plasticity, damage mechanics, and ANNs resulted in an accurate prediction of fatigue life and crack initiation points under complex loading conditions of single crystals superalloys. Full article
Show Figures

Figure 1

Figure 1
<p>Dimensions of the notched specimen.</p>
Full article ">Figure 2
<p>Octahedral (<b>a</b>) and cubic (<b>b</b>) slip planes of a FCC crystal.</p>
Full article ">Figure 3
<p>Demonstration of opening modes of a crack.</p>
Full article ">Figure 4
<p>Structure of artificial neural networks designed for regression tasks, demonstrating input, hidden, and output layers.</p>
Full article ">Figure 5
<p>Finite element mesh (<b>a</b>) and the boundary conditions (<b>b</b>) of the notched specimen.</p>
Full article ">Figure 6
<p>Stress in 22-direction (<b>a</b>), logarithmic strain in 22-direction (<b>b</b>), and lifetime (<b>c</b>) contours of a single crystal notch specimen under HCF with R = 0.6 and cut orientation of (100)[001].</p>
Full article ">Figure 7
<p>Fatigue crack initiation at the surface of SC16 nickel-base single crystal with tensile crystal orientation &lt;001&gt; is represented in (<b>a</b>). Reproduced with permission from S. Forest, Crystal plasticity and damage at cracks and notches in nickel-base single-crystal superalloys; published by Elsevier, 2022 [<a href="#B33-applsci-15-00201" class="html-bibr">33</a>]. Lifetime assessment of the same specimen orientations is represented in (<b>b</b>), where the minimum lifetime on the FEA model is marked as the predicted crack initiation location.</p>
Full article ">Figure 8
<p>Fatigue crack initiation at the surface of SC16 nickel-base single crystal with tensile crystal orientations &lt;011&gt; and &lt;111&gt; are represented in (<b>a</b>,<b>c</b>). Reproduced with permission from S. Forest, crystal plasticity and damage at cracks and notches in nickel-base single-crystal superalloys; published by Elsevier, 2022 [<a href="#B33-applsci-15-00201" class="html-bibr">33</a>]. Lifetime assessment of the same specimen orientations is represented in (<b>b</b>,<b>d</b>), respectively, in which the minimum lifetime on the FEA model is marked as the predicted crack initiation location.</p>
Full article ">Figure 8 Cont.
<p>Fatigue crack initiation at the surface of SC16 nickel-base single crystal with tensile crystal orientations &lt;011&gt; and &lt;111&gt; are represented in (<b>a</b>,<b>c</b>). Reproduced with permission from S. Forest, crystal plasticity and damage at cracks and notches in nickel-base single-crystal superalloys; published by Elsevier, 2022 [<a href="#B33-applsci-15-00201" class="html-bibr">33</a>]. Lifetime assessment of the same specimen orientations is represented in (<b>b</b>,<b>d</b>), respectively, in which the minimum lifetime on the FEA model is marked as the predicted crack initiation location.</p>
Full article ">Figure 9
<p>Comparison of numerical and experimental results of CMSX4 notched specimen under HCF loading at 750 °C with R <math display="inline"><semantics> <mrow> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (<b>a</b>) and R <math display="inline"><semantics> <mrow> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math> (<b>b</b>).</p>
Full article ">Figure 10
<p>Comparison of modified Chaboche model (CM) and experimental results of CMSX4 at 600 °C for R = 0 and R = 0.4 (<b>a</b>) and at 900 °C for R = 0 (<b>b</b>).</p>
Full article ">Figure 11
<p>Fatigue curves produced on training data spectrum for different temperatures and R ratios and their comparison with the experimental data (<b>a</b>) and for different R ratios at the same temperature (<b>b</b>).</p>
Full article ">Figure 12
<p>Fatigue curves prediction at various temperatures for CMSX-4.</p>
Full article ">
22 pages, 2379 KiB  
Article
Harnessing Convolutional Neural Networks for Automated Wind Turbine Blade Defect Detection
by Mislav Spajić, Mirko Talajić and Mirjana Pejić Bach
Designs 2025, 9(1), 2; https://doi.org/10.3390/designs9010002 - 27 Dec 2024
Viewed by 408
Abstract
The shift towards renewable energy, particularly wind energy, is rapidly advancing globally, with Southeastern Europe and Croatia, in particular, experiencing a notable increase in wind turbine construction. The frequent exposure of wind turbine blades to environmental stressors and operational forces requires regular inspections [...] Read more.
The shift towards renewable energy, particularly wind energy, is rapidly advancing globally, with Southeastern Europe and Croatia, in particular, experiencing a notable increase in wind turbine construction. The frequent exposure of wind turbine blades to environmental stressors and operational forces requires regular inspections to identify defects, such as erosion, cracks, and lightning damage, in order to minimize maintenance costs and operational downtime. This study aims to develop a machine learning model using convolutional neural networks to simplify the defect detection process for wind turbine blades, enhancing the efficiency and accuracy of inspections conducted by drones. The model leverages transfer learning on the YOLOv7 architecture and is trained on a dataset of 231 images with 246 annotated defects across eight categories, achieving a mean average precision of 0.76 at an intersection over the union threshold of 0.5. This research not only presents a robust framework for automated defect detection but also proposes a methodological approach for future studies in deep learning for structural inspections, highlighting significant economic benefits and improvements in inspection quality and speed. Full article
(This article belongs to the Special Issue Design and Analysis of Offshore Wind Turbines)
Show Figures

Figure 1

Figure 1
<p>Examples of four types of defects on blades [<a href="#B11-designs-09-00002" class="html-bibr">11</a>], adapted by the authors.</p>
Full article ">Figure 2
<p>Cropping region of interest and resizing image to suitable dimensions in cropall.</p>
Full article ">Figure 3
<p>Labeling a defect in LabelImg; Erosion defect with its corresponding bounding box.</p>
Full article ">Figure 4
<p>Examples of images without any augmentations and with labeled defects.</p>
Full article ">Figure 5
<p>Result of mosaic augmentation for the purpose of artificially enlarging input data.</p>
Full article ">Figure 6
<p>Precision–Recall curve per class.</p>
Full article ">Figure 7
<p>Progression of the cost function results through 400 epochs of training, which indicates no overfit.</p>
Full article ">Figure 8
<p>Methodological framework for future research.</p>
Full article ">
20 pages, 6539 KiB  
Article
Adaptive Neuro-Fuzzy System for Detection of Wind Turbine Blade Defects
by Lesia Dubchak, Anatoliy Sachenko, Yevgeniy Bodyanskiy, Carsten Wolff, Nadiia Vasylkiv, Ruslan Brukhanskyi and Volodymyr Kochan
Energies 2024, 17(24), 6456; https://doi.org/10.3390/en17246456 - 21 Dec 2024
Viewed by 522
Abstract
Wind turbines are the most frequently used objects of renewable energy today. However, issues that arise during their operation can greatly affect their effectiveness. Blade erosion, cracks, and other defects can slash turbine performance while also forcing maintenance costs to soar. Modern defect [...] Read more.
Wind turbines are the most frequently used objects of renewable energy today. However, issues that arise during their operation can greatly affect their effectiveness. Blade erosion, cracks, and other defects can slash turbine performance while also forcing maintenance costs to soar. Modern defect detection applications have significant computing resources needed for training and insufficient accuracy. The goal of this study is to develop the improved adaptive neuro-fuzzy inference system (ANFIS) for wind turbine defect detection, which will reduce computing resources and increase its accuracy. Unmanned aerial vehicles are deployed to photograph the turbines, and these images are beamed back and processed for early defect detection. The proposed adaptive neuro-fuzzy inference system processes the data vectors with lower complexity and higher accuracy. For this purpose, the authors explored grid partitioning and subtractive clustering methods and selected the last one because it uses three rules only for fault detection, ensuring low computational costs and enabling the discovery of wind turbine defects quickly and efficiently. Moreover, the proposed ANFIS is implemented in a controller, which has an accuracy of 91%, that is 1.4 higher than the accuracy of the existing similar controller. Full article
(This article belongs to the Special Issue Smart Optimization and Renewable Integrated Energy System)
Show Figures

Figure 1

Figure 1
<p>Typical components of the adaptive neuro-fuzzy system.</p>
Full article ">Figure 2
<p>Architecture of the adaptive neuro-fuzzy system ([<a href="#B44-energies-17-06456" class="html-bibr">44</a>], reworked by authors).</p>
Full article ">Figure 3
<p>An example of an image for a wind turbine with corrosion [<a href="#B45-energies-17-06456" class="html-bibr">45</a>].</p>
Full article ">Figure 4
<p>General scheme of a fuzzy system generated by grid partitioning.</p>
Full article ">Figure 5
<p>Fragment of the rule base for the fuzzy wind turbine blade defect detection system (generated by the grid partitioning).</p>
Full article ">Figure 6
<p>General scheme of the fuzzy system generated by subtractive clustering.</p>
Full article ">Figure 7
<p>Demonstration of the fuzzy system generated by subtractive clustering.</p>
Full article ">Figure 8
<p>The structure of the ANFIS for wind turbine blade defects detection generated by the grid partitioning.</p>
Full article ">Figure 9
<p>Structure of ANFIS for blade defects detection generated by the subtractive clustering.</p>
Full article ">Figure 10
<p>The process of training a neuro-fuzzy system based on a fuzzy system generated by the grid partitioning.</p>
Full article ">Figure 11
<p>The process of training a neuro-fuzzy system based on a fuzzy system generated by subtractive clustering.</p>
Full article ">Figure 12
<p>The result of testing a neuro-fuzzy system based on grid partitioning (<span style="color:#00B0F0">o</span>—training data, <span style="color:red">*</span>—FIS output).</p>
Full article ">Figure 13
<p>The result of testing a neuro-fuzzy system based on subtractive clustering (<span style="color:#00B0F0">o</span>—training data, <span style="color:red">*</span>—FIS output).</p>
Full article ">Figure 14
<p>General scheme of the neuro-fuzzy controller for blade defect detection of wind turbine.</p>
Full article ">Figure 15
<p>Display of the values for the input variables of the neuro-fuzzy controller.</p>
Full article ">Figure 16
<p>The value of the output variable for the neuro-fuzzy controller.</p>
Full article ">
13 pages, 1955 KiB  
Article
Numerical Study on the Static Bending Response of Cracked Wind Turbine Blades Reinforced with Graphene Platelets
by Hyeong Jin Kim and Jin-Rae Cho
Nanomaterials 2024, 14(24), 2020; https://doi.org/10.3390/nano14242020 - 16 Dec 2024
Viewed by 536
Abstract
With the growing demand for wind energy, the development of advanced materials for wind turbine support structures and blades has garnered significant attention in both industry and academia. In previous research, the authors investigated the incorporation of graphene platelets (GPLs) into wind turbine [...] Read more.
With the growing demand for wind energy, the development of advanced materials for wind turbine support structures and blades has garnered significant attention in both industry and academia. In previous research, the authors investigated the incorporation of graphene platelets (GPLs) into wind turbine blades, focusing on the structural performance and cost-effectiveness relative to conventional fiberglass composites. These studies successfully demonstrated the potential advantages of GPL reinforcement in improving blade performance and reducing the blade’s weight and costs. Building upon prior work, the present study conducts a detailed investigation into the static bending behavior of GPL-reinforced wind turbine blades, specifically examining the impact of crack location and length. A finite element model of the SNL 61.5 m wind turbine blade was rigorously developed and validated through comparison with the existing literature to ensure its accuracy. Comprehensive parametric analyses were performed to assess deflection under various crack lengths and positions, considering both flapwise and edgewise bending deformations. The findings indicate that GPL-reinforced blades exhibit reduced sensitivity to crack propagation compared to traditional fiberglass blades. Furthermore, the paper presents a thorough parametric analysis of the effects of crack location and length on blade performance. Full article
Show Figures

Figure 1

Figure 1
<p>The SNL 61.5 m wind turbine blade model: (<b>a</b>) finite element model and (<b>b</b>) loading and boundary conditions.</p>
Full article ">Figure 2
<p>Aerodynamic load distribution calculated using BEMT: (<b>a</b>) normal and tangential forces, (<b>b</b>) pitching moment.</p>
Full article ">Figure 3
<p>Location of cracks in the wind turbine blade.</p>
Full article ">Figure 4
<p>Composing parts of wind turbine blade.</p>
Full article ">Figure 5
<p>Blade deflection along the blade span [<a href="#B38-nanomaterials-14-02020" class="html-bibr">38</a>,<a href="#B39-nanomaterials-14-02020" class="html-bibr">39</a>,<a href="#B40-nanomaterials-14-02020" class="html-bibr">40</a>]: (<b>a</b>) flapwise deflection and (<b>b</b>) edgewise deflection.</p>
Full article ">Figure 6
<p>Maximum edgewise deflection in the wind turbine blade with cracks in the LP.</p>
Full article ">Figure 7
<p>Stress distribution in the GPL-reinforced wind turbine blade with the worst-case crack in the LP.</p>
Full article ">Figure 8
<p>Maximum edgewise deflection in the wind turbine blade with cracks in the TP.</p>
Full article ">Figure 9
<p>Stress distribution in the GPL-reinforced wind turbine blade with the worst-case crack in the TP.</p>
Full article ">Figure 10
<p>Maximum flapwise deflection in a GPL-reinforced wind turbine blade with cracks on the spar cap and shear web.</p>
Full article ">Figure 11
<p>Stress distribution in the GPL-reinforced wind turbine blade with the worst-case crack in the spar cap.</p>
Full article ">Figure 12
<p>Increase in maximum deflection of wind turbine blades with cracks in different parts.</p>
Full article ">
21 pages, 6275 KiB  
Article
Design Optimization of a Marine Propeller Shaft for Enhanced Fatigue Life: An Integrated Computational Approach
by Víctor Tuninetti, Diego Martínez, Sunny Narayan, Brahim Menacer and Angelo Oñate
J. Mar. Sci. Eng. 2024, 12(12), 2227; https://doi.org/10.3390/jmse12122227 - 5 Dec 2024
Viewed by 699
Abstract
This study investigates the design and potential failure modes of a marine propeller shaft using computational and analytical methods. The aim is to assess the structural integrity of the existing design and propose modifications for improved reliability and service life. Analytical calculations based [...] Read more.
This study investigates the design and potential failure modes of a marine propeller shaft using computational and analytical methods. The aim is to assess the structural integrity of the existing design and propose modifications for improved reliability and service life. Analytical calculations based on classification society rules determined acceptable shaft diameter ranges, considering torsional shear stress limits for SAE 1030 steel. A Campbell diagram analysis identified potential resonance issues at propeller blade excitation frequencies, leading to a recommended operating speed reduction for a safety margin. Support spacing was determined using both the Ship Vibration Design Guide and an empirical method, with the former yielding more conservative results. Finite element analysis, focusing on the keyway area, revealed stress concentrations approaching the material’s ultimate strength. A mesh sensitivity analysis ensured accurate stress predictions. A round-ended rectangular key geometry modification showed a significant stress reduction. Fatigue life analysis using the Goodman equation, incorporating various factors, predicted infinite life under different loading conditions, but varying safety factors highlighted the impact of these conditions. The FEA revealed that the original keyway design led to stress concentrations exceeding allowable limits, correlating with potential shaft failure. The proposed round-ended rectangular key geometry significantly reduced stress, mitigating the risk of fatigue crack initiation. This research contributes to the development of more reliable marine propulsion systems by demonstrating the efficacy of integrating analytical methods, finite element simulations, and fatigue life predictions in the design process. Full article
(This article belongs to the Section Ocean Engineering)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the propulsion system with structural details of (<b>b</b>) shaft, (<b>c</b>) keys, (<b>d</b>) propeller, and couplings.</p>
Full article ">Figure 2
<p>(<b>a</b>) Propeller shaft showing (<b>b</b>) delamination in the propeller side keyseat.</p>
Full article ">Figure 3
<p>Classification of loads producing fatigue; (<b>a</b>) alternating stress; (<b>b</b>) fluctuating stress.</p>
Full article ">Figure 4
<p>Different meshing configurations for the sensitivity analysis.</p>
Full article ">Figure 5
<p>Campbell diagram of the propeller shaft system, showing the intersection of the 3× propeller excitation frequency with the first natural frequency of the shaft.</p>
Full article ">Figure 6
<p>(<b>a</b>) Marine propeller shaft model in Shaftdesigner, showing deformation directions (arrows) corresponding to vibration modes. (<b>b</b>) Bending behavior under dynamic loads at a safe operating speed of 442 rpm. (<b>c</b>) Dynamic torsional load variation with rotational speed.</p>
Full article ">Figure 7
<p>Loading cycles of the propeller shaft for three cases: (<b>a</b>) continuous operation at a constant recommended working speed, (<b>b</b>) stationary and operating moments, and (<b>c</b>) forward and reverse operation.</p>
Full article ">Figure 8
<p>Mesh sensitivity analysis.</p>
Full article ">Figure 9
<p>Simulation results of the propeller shaft. (<b>a</b>) Overall view of equivalent stresses. (<b>b</b>) Local stresses distribution in keyways. Black box showing maximum local values.</p>
Full article ">Figure 10
<p>Localized stress distribution in the proposed redesigned keyway channels of the marine propeller shaft. Dimensions: (<b>a</b>) propeller side key—32 mm × 18 mm × 250 mm; (<b>b</b>) coupling side key—30 mm × 18 mm × 160 mm.</p>
Full article ">
20 pages, 10274 KiB  
Article
High-Cycle Fatigue Fracture Behavior and Stress Prediction of Ni-Based Single-Crystal Superalloy with Film Cooling Hole Drilled Using Femtosecond Laser
by Zhen Li, Yuanming Xu, Xinling Liu, Changkui Liu and Chunhu Tao
Metals 2024, 14(12), 1354; https://doi.org/10.3390/met14121354 - 27 Nov 2024
Viewed by 526
Abstract
A high-temperature, high-cycle fatigue test was conducted on a nickel-based single-crystal superalloy with a pore structure. Optical and scanning electron microscopy were utilized to examine the crack propagation paths and fatigue fracture surfaces at the macro and micro scales. The analysis of crack [...] Read more.
A high-temperature, high-cycle fatigue test was conducted on a nickel-based single-crystal superalloy with a pore structure. Optical and scanning electron microscopy were utilized to examine the crack propagation paths and fatigue fracture surfaces at the macro and micro scales. The analysis of crack initiation and propagation related to the pore structure facilitated the development of a crack shape factor reflecting these distinct fracture behaviors. Predictions about the high-cycle fatigue stress experienced by the specimen were made, accompanied by an error analysis, providing critical insights for precise stress calculations and structural optimization in engine blade design. The results reveal that high-cycle fatigue cracks originate from corner cracks at pore edges, with the initial propagation displaying smooth crystallographic plane features. Subsequent stages show clear fatigue arc patterns in the propagation zones. The fracture surface exhibits the significant layering of oxide layers, primarily composed of NiO, with traces of CoO displaying columnar growth. AL2O3 is predominantly found at the interfaces between the matrix and oxide layers. Short and straight dislocations near the oxide layers and within the matrix suggest that dislocation multiplication and planar slip dominate the slip mechanisms in this alloy. The orientation of the fracture surface is mainly perpendicular to the load direction, with minor inclined facets in localized areas. Correlations were established between the plastic zone dimensions at the crack tips and the corresponding fatigue stresses. Without grain boundaries in single-crystal alloys, these dimensions are easily derived as parameters for fatigue stress analysis. The selected crack shape factor, “elliptical corner crack at pore edges”, captures the initiation and propagation traits relevant to porous structures. Subsequent calculations, accounting for the impact of oxide layers on stress assessments, indicated an error ratio ranging from 1.00 to 1.21 compared to nominal stress values. Full article
Show Figures

Figure 1

Figure 1
<p>Microstructural characteristics of enhanced phase (γ′-Ni<sub>3</sub>Al) morphology.</p>
Full article ">Figure 2
<p>(<b>a</b>) Structural morphology of pore walls in aerogel; (<b>b</b>) morphological characteristics of pore edge structure.</p>
Full article ">Figure 3
<p>(<b>a</b>) Geometrical dimensions of high-cycle fatigue specimens (all dimensions are in mm); (<b>b</b>) schematic diagram of the vibration excitation system; (<b>c</b>) monitoring position (red dot) and strain gauge position (yellow square) of the laser displacement sensor.</p>
Full article ">Figure 4
<p>High-cycle fatigue life of specimens with perforations compared to those without perforations (specimens marked in yellow for stress prediction analysis).</p>
Full article ">Figure 5
<p>(<b>a</b>) Microstructural characteristics of cross-section of M9; (<b>b</b>) Microstructural characteristics of cross-section of M1; (<b>c</b>) Microstructural characteristics of cross-section of M4.</p>
Full article ">Figure 6
<p>(<b>a</b>) Microstructural characteristics of the cross-section of M9; (<b>b</b>) Microstructural characteristics of the cross-section of M1; (<b>c</b>) Microstructural characteristics of the cross-section of M4.</p>
Full article ">Figure 7
<p>(<b>a</b>) Analysis of the morphologies of oxidized particles (The yellow box shows the EDS analysis area); (<b>b</b>) Euler angle results of oxidized particles (Different colors represent different grain orientations).</p>
Full article ">Figure 8
<p>(<b>a</b>) Back to bottom plot of HAADF elemental analysis; (<b>b</b>) Elemental distribution map.</p>
Full article ">Figure 9
<p>Characteristics of dislocation morphology.</p>
Full article ">Figure 10
<p>Schematic representation of the plastic deformation zone at the crack tip.</p>
Full article ">Figure 11
<p>Schematic representation of the EBSD sample preparation procedure.</p>
Full article ">Figure 12
<p>Results and data analysis of the KAM experiment.</p>
Full article ">Figure 13
<p>(<b>a</b>) Schematic representation of the elliptical corner crack model; (<b>b</b>) crack extension direction versus θ angle definition plot; (<b>c</b>) the cross-section of a high-cycle simulation specimen.</p>
Full article ">Figure 14
<p>Meshing and convergence study of M4 specimens.</p>
Full article ">Figure 15
<p>Finite element analysis of stress distribution at the chip placement interface of the M4 specimen (red marked positions).</p>
Full article ">Figure 16
<p>Stress along the crack propagation path for the M4 specimen.</p>
Full article ">
20 pages, 8467 KiB  
Article
Quantitative Detection Method for Surface Angled Cracks Based on Laser Ultrasonic Full-Field Scanning Data
by Chenwei Wang, Rui Han, Yihui Zhang, Yuzhong Wang, Yanyang Zi and Jiyuan Zhao
Sensors 2024, 24(23), 7519; https://doi.org/10.3390/s24237519 - 25 Nov 2024
Viewed by 472
Abstract
Surface angled cracks on critical components in high-speed machinery can lead to fractures under stress and pressure, posing a significant threat to the operational safety of equipment. To detect surface angled cracks on critical components, this paper proposes a “Quantitative Detection Method for [...] Read more.
Surface angled cracks on critical components in high-speed machinery can lead to fractures under stress and pressure, posing a significant threat to the operational safety of equipment. To detect surface angled cracks on critical components, this paper proposes a “Quantitative Detection Method for Surface Angled Cracks Based on Full-field Scanning Data”. By analyzing different ultrasonic signals in the full-field scanning data from laser ultrasonics, the width, angle, and length of surface angled cracks can be determined. This study investigates the propagation behavior of ultrasonic waves and their interaction with surface angled cracks through theoretical calculations. The crack width is solved by analyzing the distribution of Rayleigh waves in the full-field scanning data. This paper also discusses the differences in ultrasonic wave propagation between near-field and far-field detection and identifies the critical point between these regions. Different computational methods are employed to calculate the inclination angle and the crack endpoint at various scan positions. Four sets of experiments were conducted to validate the proposed method, with results showing that the errors in determining the width, angle, and length of the surface angled cracks were all within 5%. This confirms the feasibility of the method for detecting surface angled cracks. The quantitative detection of surface angled cracks on critical components using this method allows for a comprehensive assessment of the component’s condition, aiding in the prediction of service life and the mitigation of operational risks. This method shows promising application potential in areas such as aircraft engine blade inspection and gear inspection. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Figure 1

Figure 1
<p>The core concept of Quantitative Detection Method for Surface Angled Cracks Based on Full-Field Scanning Data.</p>
Full article ">Figure 2
<p>Schematic diagram of surface angled crack full-field scanning and full-field scanning signals for surface angled cracks. (<b>a</b>) Schematic diagram of surface angled crack full-field scanning. (<b>b</b>) Full-field scanning signals for surface angled cracks.</p>
Full article ">Figure 3
<p>Schematic diagram and abstract geometric model for near-field surface angled crack detection. (<b>a</b>) Schematic diagram, (<b>b</b>) Abstract geometric model.</p>
Full article ">Figure 4
<p>Schematic diagram and abstract geometric model for far-field surface angled crack detection. (<b>a</b>) Schematic diagram, (<b>b</b>) Abstract geometric model.</p>
Full article ">Figure 5
<p>(<b>a</b>) Diagram between generation position and calculated inclination angle; (<b>b</b>) Elliptical positioning algorithm for locating crack endpoint in far-field detection. (<b>a</b>) Calculated inclination angle. (<b>b</b>) Result of far-field detection.</p>
Full article ">Figure 6
<p>Laser ultrasonic testing system.</p>
Full article ">Figure 7
<p>Specimens and schematic diagram of experiments. (<b>a</b>) Specimens, (<b>b</b>) Schematic diagram of experiments.</p>
Full article ">Figure 8
<p>Raw full-field b-scan signals for surface angled cracks with different inclinations. (<b>a</b>) Experiment 1#, (<b>b</b>) Experiment 2#, (<b>c</b>) Experiment 3#, (<b>d</b>) Experiment 4#.</p>
Full article ">Figure 9
<p>Raw signal for experiment 3#. (The red dashed line represents the time window).</p>
Full article ">Figure 10
<p>Extraction of ultrasonic propagation time from the signal: (<b>a</b>) Raw signal; (<b>b</b>) Defect-free reference signal; (<b>c</b>) Result of subtracting the reference signal from the raw signal; (<b>d</b>) Extraction of the longitudinal wave defect echo; (<b>e</b>) Extraction of the Rayleigh wave defect echo.</p>
Full article ">Figure 11
<p>Diagram between generation position and calculated inclination angle. (<b>a</b>) Experiment 1#, (<b>b</b>) Experiment 2#, (<b>c</b>) Experiment 3#, (<b>d</b>) Experiment 4#.</p>
Full article ">Figure 12
<p>Elliptical positioning algorithm for locating endpoints of far-field surface angled cracks. (<b>a</b>) Experiment 1#, (<b>b</b>) Experiment 1#, (<b>c</b>) Experiment 2#, (<b>d</b>) Experiment 2#, (<b>e</b>) Experiment 3#, (<b>f</b>) Experiment 3#, (<b>g</b>) Experiment 4#, (<b>h</b>) Experiment 4#.</p>
Full article ">Figure 12 Cont.
<p>Elliptical positioning algorithm for locating endpoints of far-field surface angled cracks. (<b>a</b>) Experiment 1#, (<b>b</b>) Experiment 1#, (<b>c</b>) Experiment 2#, (<b>d</b>) Experiment 2#, (<b>e</b>) Experiment 3#, (<b>f</b>) Experiment 3#, (<b>g</b>) Experiment 4#, (<b>h</b>) Experiment 4#.</p>
Full article ">
27 pages, 2633 KiB  
Article
Classification Analytics for Wind Turbine Blade Faults: Integrated Signal Analysis and Machine Learning Approach
by Waqar Ali, Idriss El-Thalji, Knut Erik Teigen Giljarhus and Andreas Delimitis
Energies 2024, 17(23), 5856; https://doi.org/10.3390/en17235856 - 22 Nov 2024
Viewed by 565
Abstract
Wind turbine blades are critical components of wind energy systems, and their structural health is essential for reliable operation and maintenance. Several studies have used time-domain and frequency-domain features alongside machine learning techniques to predict faults in wind turbine blades, such as erosion [...] Read more.
Wind turbine blades are critical components of wind energy systems, and their structural health is essential for reliable operation and maintenance. Several studies have used time-domain and frequency-domain features alongside machine learning techniques to predict faults in wind turbine blades, such as erosion and cracks. However, a key gap remains in integrating these methods into a unified framework for fault prediction, which could offer a more comprehensive solution for diagnosing faults. This paper presents an approach to classify faults in wind turbine blades by leveraging well-known signals and analysis with machine learning techniques. The methodology involves a detailed feature engineering process that extracts and analyzes features from the time and frequency domains. Open-source vibration data collected from an experimental setup (where a small wind turbine with an artificially eroded and cracked blade was tested) were utilized. The time- and frequency-domain features were extracted and analyzed using various machine learning algorithms. It was found that erosion and crack faults have unique time- and frequency-domain features. The crack fault introduces an amplitude modulation in the vibration time wave, which produces sidebands around the fundamental frequency in the frequency domain. However, erosion fault introduces asymmetricity and flatness to the vibration time wave, which produces harmonics in the frequency-domain plot. The results also highlighted that utilizing both time- and frequency-fault features enhances the performance of the machine learning algorithms. This study further illustrates that even though some machine learning algorithms provide similar high classification accuracy, they might differ in quantifying error Types I, II, and, III, which is extremely important for maintenance engineers, as it might lead to undetected fault events and false alarm events. Full article
(This article belongs to the Special Issue Structural Testing and Health Monitoring of Wind Turbines)
Show Figures

Figure 1

Figure 1
<p>Overview of the methodology used in this study, illustrating key components and workflows for data processing and analysis.</p>
Full article ">Figure 2
<p>Time series signals with overlapping segments highlighted. The plot shows the entire signal in blue, with the first two overlapping segments highlighted in black and dark red, respectively. Each segment is annotated with its sample number.</p>
Full article ">Figure 3
<p>Confusion matrix visualization showing Type I, Type II, and Type III errors for predicted crack, erosion, and healthy label in relation to the true label.</p>
Full article ">Figure 4
<p>The raw signals and their autocorrelation responses for wind turbine blade faults: healthy, crack, and erosion conditions.</p>
Full article ">Figure 5
<p>Comparison of normalized time-domain features across fault types.</p>
Full article ">Figure 6
<p>FFT spectrum for healthy, crack fault, and erosion fault blade conditions.</p>
Full article ">Figure 7
<p>(a) XGBoost feature importance, (<b>b</b>) XGBoost confusion matrix.</p>
Full article ">Figure 8
<p>(<b>a</b>) Random forest confusion matrix, (<b>b</b>) support vector machine confusion matrix.</p>
Full article ">Figure 9
<p>(<b>a</b>) Logistic regression confusion matrix, (<b>b</b>) MLP classifier confusion matrix.</p>
Full article ">Figure 10
<p>Comparison between Type I, Type II, and Type III errors for the models (time domain, frequency domain, and combined features). (<b>a</b>) Error quantification based on time-domain features. (<b>b</b>) Error quantification based on frequency-domain features. (<b>c</b>) Error quantification based on combined features.</p>
Full article ">
15 pages, 8736 KiB  
Article
Research on Classification and Identification of Crack Faults in Steam Turbine Blades Based on Supervised Contrastive Learning
by Qinglei Zhang, Laifeng Tang, Jiyun Qin, Jianguo Duan and Ying Zhou
Entropy 2024, 26(11), 956; https://doi.org/10.3390/e26110956 - 6 Nov 2024
Viewed by 633
Abstract
Steam turbine blades may crack, break, or suffer other failures due to high temperatures, high pressures, and high-speed rotation, which seriously threatens the safety and reliability of the equipment. The signal characteristics of different fault types are slightly different, making it difficult to [...] Read more.
Steam turbine blades may crack, break, or suffer other failures due to high temperatures, high pressures, and high-speed rotation, which seriously threatens the safety and reliability of the equipment. The signal characteristics of different fault types are slightly different, making it difficult to accurately classify the faults of rotating blades directly through vibration signals. This method combines a one-dimensional convolutional neural network (1DCNN) and a channel attention mechanism (CAM). 1DCNN can effectively extract local features of time series data, while CAM assigns different weights to each channel to highlight key features. To further enhance the efficacy of feature extraction and classification accuracy, a projection head is introduced in this paper to systematically map all sample features into a normalized space, thereby improving the model’s capacity to distinguish between distinct fault types. Finally, through the optimization of a supervised contrastive learning (SCL) strategy, the model can better capture the subtle differences between different fault types. Experimental results show that the proposed method has an accuracy of 99.61%, 97.48%, and 96.22% in the classification task of multiple crack fault types at three speeds, which is significantly better than Multilayer Perceptron (MLP), Residual Network (ResNet), Momentum Contrast (MoCo), and Transformer methods. Full article
(This article belongs to the Section Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>System framework diagram (The figure illustrates a fault diagnosis framework based on vibration signal data. The experiment collects vibration signals of blade crack faults, constructs a fault dataset, and adds Gaussian noise. The 1D CNN is combined with CAM to extract fault features. A projection head is introduced to map all sample features into a normalized space, thereby enhancing the model’s ability to distinguish between different fault types. Additionally, contrast loss and cross-entropy loss are calculated through supervised contrast learning to complete the fault classification.).</p>
Full article ">Figure 2
<p>Vibration signals before and after adding Gaussian noise. ((<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) represent the “normal” without adding Gaussian noise, the fault signal of “crack1”, the fault signal of “crack2”, the fault signal of “crack3”, and the fault signal of “fracture”, respectively; (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) represent the above five signals after adding Gaussian noise, respectively.).</p>
Full article ">Figure 3
<p>Dynamic test bench of rotor system with integral shroud blade.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) represent the accuracy of MLP, ResNet, MoCo, Transformer, and our proposed method at 1400 r/min, 1800 r/min, and 2200 r/min, respectively.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) represent the confusion matrix results at 1400 r/min, 1800 r/min, and 2200 r/min. The values on the diagonal represent the number of samples predicted correctly, and the values on the off-diagonal represent the number of samples predicted incorrectly.</p>
Full article ">Figure 6
<p>t-SNE feature visualization: (<b>a</b>) 1400 r/min, (<b>b</b>) 1800 r/min, (<b>c</b>) 2200 r/min.</p>
Full article ">
22 pages, 9204 KiB  
Article
Analysis of the Nonlinear Complex Response of Cracked Blades at Variable Rotational Speeds
by Bo Shao, Chenguang Fan, Shunguo Fu and Jin Zeng
Machines 2024, 12(10), 725; https://doi.org/10.3390/machines12100725 - 14 Oct 2024
Viewed by 854
Abstract
The operation of an aero-engine involves various non-stationary processes of acceleration and deceleration, with rotational speed varying in response to changing working conditions to meet different power requirements. To investigate the nonlinear dynamic behaviour of cracked blades under variable rotational speed conditions, this [...] Read more.
The operation of an aero-engine involves various non-stationary processes of acceleration and deceleration, with rotational speed varying in response to changing working conditions to meet different power requirements. To investigate the nonlinear dynamic behaviour of cracked blades under variable rotational speed conditions, this study constructed a rotating blade model with edge-penetrating cracks and proposes a component modal synthesis method that accounts for time-varying rotational speed. The nonlinear response behaviours of cracked blades were examined under three distinct operating conditions: spinless, steady speed, and non-constant speed. The findings indicated a competitive relationship between the effects of rotational speed fluctuations and unbalanced excitation on crack nonlinearity. Variations in rotational speed dominated when rotational speed perturbation was minimal; conversely, aerodynamic forces dominated when the effects of rotational speed were pronounced. An increase in rotational speed perturbation enhanced the super-harmonic nonlinearity induced by cracks, elevated the nonlinear damage index (NDI), and accentuated the crack breathing effect. As the perturbation coefficient increased, the super-harmonic nonlinearity of the crack intensified, resulting in a more complex vibration form and phase diagram. Full article
(This article belongs to the Special Issue Nonlinear Dynamics of Mechanical Systems and Machines)
Show Figures

Figure 1

Figure 1
<p>The schematic of edge penetration cracked blade.</p>
Full article ">Figure 2
<p>Schematic diagram of cracked surface contact calculation.</p>
Full article ">Figure 3
<p>FE model of the cracked blade: (<b>a</b>) edge penetration of the cracked blade, (<b>b</b>) progressive meshing of the cracked region, (<b>c</b>) top view of the blade.</p>
Full article ">Figure 4
<p>Reduction flow of the CMS–Solid, accounting for time-varying rotational speeds.</p>
Full article ">Figure 5
<p>Effects of <span class="html-italic">a</span> on the first six-order frequencies convergence: (<b>a</b>) CMS, (<b>b</b>) CMS–Solid.</p>
Full article ">Figure 6
<p><span class="html-italic">x</span>-direction displacement: (<b>a</b>) FE model, (<b>b</b>) CMS model, and (<b>c</b>) CMS–Solid model.</p>
Full article ">Figure 7
<p><span class="html-italic">x</span>-direction displacement difference with FE model: (<b>a</b>) CMS model, (<b>b</b>) CMS–Solid model.</p>
Full article ">Figure 8
<p>Comparison of dynamic responses for three calculation methods at a steady rotational speed: (<b>a</b>) steady-state time-domain response results, (<b>b</b>) frequency-domain results.</p>
Full article ">Figure 9
<p>Calculation time for transient response across three calculation methods.</p>
Full article ">Figure 10
<p>Comparison of dynamic responses between FEM and CMS–Solid method at unsteady rotational speed: (<b>a</b>) overall response curve, (<b>b</b>) response magnification, and (<b>c</b>) response magnification.</p>
Full article ">Figure 11
<p>Comparison of frequency domains between FEM and CMS–Solid method at unsteady rotational speeds.</p>
Full article ">Figure 12
<p><span class="html-italic">x</span>-direction response of the cracked blade tip for different amplitudes of rotational speed disturbance: (<b>a</b>) k = 0, (<b>b</b>) k = 2%, (<b>c</b>) k = 4%, (<b>d</b>) k = 6%, (<b>e</b>) k = 8%, and (<b>f</b>) k = 10%.</p>
Full article ">Figure 13
<p>Enlarged <span class="html-italic">x</span>-direction response of the cracked blade tip for different amplitudes of rotational speed disturbance: (<b>a</b>) k = 0, (<b>b</b>) k = 2%, (<b>c</b>) k = 4%, (<b>d</b>) k = 6%, (<b>e</b>) k = 8%, and (<b>f</b>) k = 10%.</p>
Full article ">Figure 14
<p>Frequency spectrum of the cracked blade for different speed disturbance amplitudes: (<b>a</b>) k = 0, (<b>b</b>) k = 2%, (<b>c</b>) k = 4%, (<b>d</b>) k = 6%, (<b>e</b>) k = 8%, and (<b>f</b>) k = 10%.</p>
Full article ">Figure 15
<p>Phase diagram of cracked blade at different speed disturbance amplitudes: (<b>a</b>) k = 0, (<b>b</b>) k = 2%, (<b>c</b>) k = 4%, (<b>d</b>) k = 6%, (<b>e</b>) k = 8%, and (<b>f</b>) k = 10%.</p>
Full article ">Figure 16
<p>Time-course curves of contact stresses of cracked blade at different rotational speed disturbance amplitudes: (<b>a</b>) k = 0, (<b>b</b>) k = 2%, (<b>c</b>) k = 4%, (<b>d</b>) k = 6%, (<b>e</b>) k = 8%, and (<b>f</b>) k = 10%.</p>
Full article ">Figure 17
<p><span class="html-italic">x</span>-direction response of cracked blade with varying amplitudes of rotational speed perturbation: (<b>a</b>) overall response curve, (<b>b</b>) magnified response curve.</p>
Full article ">Figure 18
<p>Frequency spectrum of a cracked blade for different speed disturbance amplitudes.</p>
Full article ">Figure 19
<p>NDI of a cracked blade under varying rotational speed disturbance amplitudes.</p>
Full article ">Figure 20
<p>Phase diagram of cracked blades at different rotational speed disturbance amplitudes.</p>
Full article ">Figure 21
<p>Time course of contact stresses on a cracked blade under different rotational speed disturbance amplitudes: (<b>a</b>) overall contact stresses and (<b>b</b>) magnified view of contact stresses.</p>
Full article ">
22 pages, 5749 KiB  
Article
DCW-YOLO: An Improved Method for Surface Damage Detection of Wind Turbine Blades
by Li Zou, Anqi Chen, Chunzi Li, Xinhua Yang and Yibo Sun
Appl. Sci. 2024, 14(19), 8763; https://doi.org/10.3390/app14198763 - 28 Sep 2024
Cited by 1 | Viewed by 1282
Abstract
Wind turbine blades (WTBs) are prone to damage from their working environment, including surface peeling and cracks. Early and effective detection of surface defects on WTBs can avoid complex and costly repairs and serious safety hazards. Traditional object detection methods have disadvantages of [...] Read more.
Wind turbine blades (WTBs) are prone to damage from their working environment, including surface peeling and cracks. Early and effective detection of surface defects on WTBs can avoid complex and costly repairs and serious safety hazards. Traditional object detection methods have disadvantages of insufficient detection capabilities, extended model inference times, low recognition accuracy for small objects, and elongated strip defects within WTB datasets. In light of these challenges, a novel model named DCW-YOLO for surface damage detection of WTBs is proposed in this research, which leverages image data collected by unmanned aerial vehicles (UAVs) and the YOLOv8 algorithm for image analysis. Firstly, Dynamic Separable Convolution (DSConv) is introduced into the C2f module of YOLOv8, allowing the model to more effectively focus on the geometric structural details associated with damage on WTBs. Secondly, the upsampling method is replaced with the content-aware reassembly of features (CARAFE), which significantly minimizes the degradation of image characteristics throughout the upsampling process and boosts the network’s ability to extract features. Finally, the loss function is substituted with the WIoU (Wise-IoU) strategy. This strategy allows for a more accurate regression of the target bounding boxes and helps to improve the reliability in the localization of WTBs damages, especially for low-quality examples. This model demonstrates a notable superiority in surface damage detection of WTBs compared to the original YOLOv8n and has achieved a substantial improvement in the [email protected] metric, rising from 91.4% to 93.8%. Furthermore, in the more rigorous [email protected]–0.95 metric, it has also seen an increase from 68.9% to 71.2%. Full article
Show Figures

Figure 1

Figure 1
<p>The structure of YOLOv8 model.</p>
Full article ">Figure 2
<p>The structure of DSConv.</p>
Full article ">Figure 3
<p>Iteration of DSConv: (<b>a</b>) Schematic of the coordinates calculation of the DSConv. (<b>b</b>) The receptive field of DSConv.</p>
Full article ">Figure 4
<p>The overall architecture of CARAFE.</p>
Full article ">Figure 5
<p>The structure of DCW-YOLO model.</p>
Full article ">Figure 6
<p>Dataset samples of WTBs: (<b>a</b>,<b>b</b>) are peeling samples. (<b>c</b>,<b>d</b>) are crack samples.</p>
Full article ">Figure 7
<p>mAP of the model: (<b>a</b>) the mAP of the original YOLOv8; (<b>b</b>) the mAP of the DCW-YOLO.</p>
Full article ">Figure 8
<p>Detection performance of the model: (<b>a</b>) the detection performance of YOLOv8; (<b>b</b>) the detection performance of DCW-YOLO.</p>
Full article ">Figure 9
<p>mAP values for different models.</p>
Full article ">Figure 10
<p>WTB damage detection results: (<b>a</b>) YOLOv5 detection effect. (<b>b</b>) YOLOv6 detection effect. (<b>c</b>) YOLOv7-Tiny detection effect. (<b>d</b>) YOLOv8 detection effect. (<b>e</b>) YOLOv10 detection effect. (<b>f</b>) DCW-YOLO detection effect.</p>
Full article ">
29 pages, 21121 KiB  
Article
Hydrodynamic Characteristics of Preloading Spiral Case and Concrete in Turbine Mode with Emphasis on Preloading Clearance
by Yutong Luo, Zonghua Li, Shaozheng Zhang, Qingfeng Ren and Zhengwei Wang
Processes 2024, 12(9), 2056; https://doi.org/10.3390/pr12092056 - 23 Sep 2024
Cited by 1 | Viewed by 686
Abstract
A pump-turbine may generate high-amplitude hydraulic excitations during operation, wherein the flow-induced response of the spiral case and concrete is a key factor affecting the stable and safe operation of the unit. The preloading spiral case can enhance the combined bearing capacity of [...] Read more.
A pump-turbine may generate high-amplitude hydraulic excitations during operation, wherein the flow-induced response of the spiral case and concrete is a key factor affecting the stable and safe operation of the unit. The preloading spiral case can enhance the combined bearing capacity of the entire structure, yet there is still limited research on the impact of the preloading pressure on the hydrodynamic response. In this study, the pressure fluctuation characteristics and dynamic behaviors of preloading a steel spiral case and concrete under different preloading pressures at rated operating conditions are analyzed based on fluid–structure interaction theory and contact model. The results show that the dominant frequency of pressure fluctuations in the spiral case is 15 fn, which is influenced by the rotor–stator interaction with a runner rotation of short and long blades. Under preloading pressures of 0.5, 0.7, and 1 times the maximum static head, higher preloading pressures reduce the contact regions, leading to uneven deformation and stress distributions with a near-positive linear correlation. The maximum deformation of the PSSC can reach 2.6 mm, and the stress is within the allowable range. The preloading pressure has little effect on the dominant frequency of the dynamic behaviors in the spiral case (15 fn), but both the maximum and amplitudes of deformation and stress increase with higher preloading pressure. The high-amplitude regions of deformation and stress along the axial direction are located near the nose vane, with maximum values of 0.003 mm and 0.082 MPa, respectively. The contact of concrete is at risk of stress concentrations and cracking under high preloading pressure. The results can provide references for optimizing the structural design and the selection of preloading pressure, which improves operation reliability. Full article
(This article belongs to the Section Process Control and Monitoring)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the work steps.</p>
Full article ">Figure 2
<p>Three-dimensional modeling and mesh of flow passage.</p>
Full article ">Figure 3
<p>The flow domain mesh of the pump-turbine unit. (<b>a</b>) Runner; (<b>b</b>) guide vanes and stay vanes.</p>
Full article ">Figure 4
<p>Fluid mesh independence analysis.</p>
Full article ">Figure 5
<p>Pressure pulsation monitoring points. (<b>a</b>) Spiral case; (<b>b</b>) inter-vane space; (<b>c</b>) zaneless space.</p>
Full article ">Figure 6
<p>Pump-turbine model test.</p>
Full article ">Figure 7
<p>Three-dimensional modeling and mesh of the preloading spiral case. (<b>a</b>) FEM model; (<b>b</b>) cross-sectional schematic of the boundary condition of the negative pressure.</p>
Full article ">Figure 8
<p>Three-dimensional modeling and mesh of the pump-turbine and plant concrete structural model.</p>
Full article ">Figure 9
<p>Structural mesh independence analysis.</p>
Full article ">Figure 10
<p>Boundary conditions of the pump-turbine unit.</p>
Full article ">Figure 11
<p>Streamline distribution of the unit. (<b>a</b>) Whole flow passage; (<b>b</b>) cross-section view.</p>
Full article ">Figure 12
<p>Pressure distribution in the flow passage.</p>
Full article ">Figure 13
<p>The time history and frequency spectra of the pressure fluctuation. (<b>a</b>) Spiral case; (<b>b</b>) inter-vane space; (<b>c</b>) vaneless space.</p>
Full article ">Figure 14
<p>The cross-sections of runner blades.</p>
Full article ">Figure 15
<p>Pressure pulsation monitoring points in S1 and S2.</p>
Full article ">Figure 16
<p>The pressure amplitude of the monitoring points on the first four dominant frequencies. (<b>a</b>) S1; (<b>b</b>) S2.</p>
Full article ">Figure 16 Cont.
<p>The pressure amplitude of the monitoring points on the first four dominant frequencies. (<b>a</b>) S1; (<b>b</b>) S2.</p>
Full article ">Figure 17
<p>Deformation and stress monitoring points (red points for total deformation, blue points for maximum principal stress).</p>
Full article ">Figure 18
<p>Contact status between the PSSC and concrete under different preloading pressures; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 18 Cont.
<p>Contact status between the PSSC and concrete under different preloading pressures; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 19
<p>The total deformation distribution of the PSSC under different preloading pressure; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 20
<p>The maximum deformation of the PSSC versus the preloading pressure.</p>
Full article ">Figure 21
<p>The time history and frequency spectra of dynamic deformation of the PSSC under different preloading pressures; (<b>a</b>) n1; (<b>b</b>) n3; (<b>c</b>) n5; (<b>d</b>) n7.</p>
Full article ">Figure 21 Cont.
<p>The time history and frequency spectra of dynamic deformation of the PSSC under different preloading pressures; (<b>a</b>) n1; (<b>b</b>) n3; (<b>c</b>) n5; (<b>d</b>) n7.</p>
Full article ">Figure 22
<p>The equivalent stress distribution of the PSSC under different preloading pressures; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 22 Cont.
<p>The equivalent stress distribution of the PSSC under different preloading pressures; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 23
<p>The time history and frequency spectra of dynamic stress of the PSSC under different preloading pressures; (<b>a</b>) n2; (<b>b</b>) n4; (<b>c</b>) n6; (<b>d</b>) n8.</p>
Full article ">Figure 23 Cont.
<p>The time history and frequency spectra of dynamic stress of the PSSC under different preloading pressures; (<b>a</b>) n2; (<b>b</b>) n4; (<b>c</b>) n6; (<b>d</b>) n8.</p>
Full article ">Figure 24
<p>The maximum principal stress distribution of the concrete under different preloading pressures; (<b>a</b>) 3.2 MPa; (<b>b</b>) 4.564 MPa; (<b>c</b>) 6.52 MPa.</p>
Full article ">Figure 25
<p>The maximum stress of the PSSC and concrete versus the preloading pressure.</p>
Full article ">
22 pages, 10776 KiB  
Article
Fatigue Characteristics Analysis of Carbon Fiber Laminates with Multiple Initial Cracks
by Zheng Liu, Yuhao Zhang, Haodong Liu, Xin Liu, Jinlong Liang and Zhenjiang Shao
Appl. Sci. 2024, 14(18), 8572; https://doi.org/10.3390/app14188572 - 23 Sep 2024
Viewed by 1234
Abstract
In the entire wind turbine system, the blade acts as the central load-bearing element, with its stability and reliability being essential for the safe and effective operation of the wind power unit. Carbon fiber, known for its high strength-to-weight ratio, high modulus, and [...] Read more.
In the entire wind turbine system, the blade acts as the central load-bearing element, with its stability and reliability being essential for the safe and effective operation of the wind power unit. Carbon fiber, known for its high strength-to-weight ratio, high modulus, and lightweight characteristics, is extensively utilized in blade manufacturing due to its superior attributes. Despite these advantages, carbon fiber composites are frequently subjected to cyclic loading, which often results in fatigue issues. The presence of internal manufacturing defects further intensifies these fatigue challenges. Considering this, the current study focuses on carbon fiber composites with multiple pre-existing cracks, conducting both static and fatigue experiments by varying the crack length, the angle between cracks, and the distance among them to understand their influence on the fatigue life under various conditions. Furthermore, this study leverages the advantages of Paris theory combined with the Extended Finite Element Method (XFEM) to simulate cracks of arbitrary shapes, introducing a fatigue simulation method for carbon fiber composite laminates with multiple cracks to analyze their fatigue characteristics. Concurrently, the Particle Swarm Optimization (PSO) algorithm is employed to determine the optimal weight configuration, and the Backpropagation neural network (BP) is used to train and adjust the weights and thresholds to minimize network errors. Building on this foundation, a surrogate model for predicting the fatigue life of carbon fiber composite laminates with multiple cracks under conditions of physical parameter uncertainty has been constructed, achieving modeling and assessment of fatigue reliability. This research offers theoretical insights and methodological guidance for the utilization of carbon fiber-reinforced composites in wind turbine blade applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the basic structure of wind turbine blades.</p>
Full article ">Figure 2
<p>Research framework diagram.</p>
Full article ">Figure 3
<p>Paris fatigue expansion curve.</p>
Full article ">Figure 4
<p>Schematic diagram of BP neural network structure.</p>
Full article ">Figure 5
<p>PSO–BP fatigue life prediction flow.</p>
Full article ">Figure 6
<p>Schematic diagram and sample of the specimen.</p>
Full article ">Figure 7
<p>Tensile test. (<b>a</b>) Test equipment. (<b>b</b>) Test process. (<b>c</b>) Failure fracture.</p>
Full article ">Figure 8
<p>Static test displacement curve and load capacity. (<b>a</b>) Maximum failure force-displacement curves of carbon fiber plates with different angles of internal cracks; (<b>b</b>) Maximum failure force and coefficient of variation of static tensile force of carbon fiber plates with different angles of internal cracks; (<b>c</b>) Maximum failure force-displacement curves of carbon fiber plates with different lengths of internal cracks; (<b>d</b>) Maximum failure force and coefficient of variation of static tensile force of carbon fiber plates with different lengths of internal cracks.</p>
Full article ">Figure 9
<p>Fatigue life with different variables. (<b>a</b>) Fatigue life of carbon fiber plates with different angles of internal cracks; (<b>b</b>) Fatigue life of carbon fiber plates with different lengths of internal cracks; (<b>c</b>) Fatigue life of carbon fiber plates with internal and edge cracks at different distances.</p>
Full article ">Figure 10
<p>Fatigue finite element model.</p>
Full article ">Figure 11
<p>Equivalent force clouds and enriched unit states for different states of carbon fiber laminates.</p>
Full article ">Figure 12
<p>Comparison of fatigue life results.</p>
Full article ">Figure 13
<p>Comparison of life prediction results of different models. (<b>a</b>) BP model; (<b>b</b>) PSO-BP model.</p>
Full article ">Figure 14
<p>Scatter plot of predicted life results.</p>
Full article ">Figure 15
<p>Probability plot of distribution of carbon fiber laminates containing multiple cracks.</p>
Full article ">Figure 16
<p>Cumulative probability of failure and reliability curves for carbon fiber laminates with multiple cracks.</p>
Full article ">Figure 17
<p>Sobol sensitivity analysis.</p>
Full article ">
21 pages, 10053 KiB  
Article
Sensitivity Analysis of Fatigue Life for Cracked Carbon-Fiber Structures Based on Surrogate Sampling and Kriging Model under Distribution Parameter Uncertainty
by Haodong Liu, Zheng Liu, Liang Tu, Jinlong Liang and Yuhao Zhang
Appl. Sci. 2024, 14(18), 8313; https://doi.org/10.3390/app14188313 - 15 Sep 2024
Viewed by 815
Abstract
The quality and reliability of wind turbine blades, as core components of wind turbines, are crucial for the operational safety of the entire system. Carbon fiber is the primary material for wind turbine blades. However, during the manufacturing process, manual intervention inevitably introduces [...] Read more.
The quality and reliability of wind turbine blades, as core components of wind turbines, are crucial for the operational safety of the entire system. Carbon fiber is the primary material for wind turbine blades. However, during the manufacturing process, manual intervention inevitably introduces minor defects, which can lead to crack propagation under complex working conditions. Due to limited understanding and measurement capabilities of the input variables of structural systems, the distribution parameters of these variables often exhibit uncertainty. Therefore, it is essential to assess the impact of distribution parameter uncertainty on the fatigue performance of carbon-fiber structures with initial cracks and quickly identify the key distribution parameters affecting their reliability through global sensitivity analysis. This paper proposes a sensitivity analysis method based on surrogate sampling and the Kriging model to address the computational challenges and engineering application difficulties in distribution parameter sensitivity analysis. First, fatigue tests were conducted on carbon-fiber structures with initial cracks to study the dispersion of their fatigue life under different initial crack lengths. Next, based on the Hashin fatigue failure criterion, a simulation analysis method for the fatigue cumulative damage life of cracked carbon-fiber structures was proposed. By introducing uncertainty parameters into the simulation model, a training sample set was obtained, and a Kriging model describing the relationship between distribution parameters and fatigue life was established. Finally, an efficient input variable sampling method using the surrogate sampling probability density function was introduced, and a Sobol sensitivity analysis method based on surrogate sampling and the Kriging model was proposed. The results show that this method significantly reduces the computational burden of distribution parameter sensitivity analysis while ensuring computational accuracy. Full article
Show Figures

Figure 1

Figure 1
<p>Manufacturing Process of Wind Turbine Blades.</p>
Full article ">Figure 2
<p>Schematic Diagram of Distribution Parameter Uncertainty Transfer.</p>
Full article ">Figure 3
<p>Sensitivity Index Solving Process Based on Kriging and Surrogate Sampling.</p>
Full article ">Figure 4
<p>Geometry of the Specimen.</p>
Full article ">Figure 5
<p>Experimental Procedure. (<b>a</b>) Tensile strength testing; (<b>b</b>) fatigue testing; (<b>c</b>) fracture details.</p>
Full article ">Figure 6
<p>Experimental Data.</p>
Full article ">Figure 7
<p>Finite Element Model Setup.</p>
Full article ">Figure 8
<p>Cumulative Fatigue Damage Flow Chart.</p>
Full article ">Figure 9
<p>Stress Cloud of Cracked Carbon Fibers.</p>
Full article ">Figure 10
<p>Comparison of Fatigue Life Simulation and Experimental Results.</p>
Full article ">Figure 11
<p>Flowchart of Cyclic Calculation.</p>
Full article ">Figure 12
<p>Model Prediction Results.</p>
Full article ">Figure 13
<p>Comparison of Life Prediction Results from Different Models.</p>
Full article ">Figure 14
<p>Fatigue Life Frequency Fitting Curves.</p>
Full article ">Figure 15
<p>Comparison of Sensitivity Index Results.</p>
Full article ">
Back to TopTop