Nothing Special   »   [go: up one dir, main page]

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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,592)

Search Parameters:
Keywords = fault location

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 5992 KiB  
Article
The First Seismic Imaging of the Holy Cross Fault in the Łysogóry Region, Poland
by Eslam Roshdy, Artur Marciniak, Rafał Szaniawski and Mariusz Majdański
Appl. Sci. 2025, 15(2), 511; https://doi.org/10.3390/app15020511 - 7 Jan 2025
Abstract
The Holy Cross Mountains represent an isolated outcrop of Palaeozoic rocks located in the Trans-European Suture Zone, which is the boundary between the Precambrian East European Craton and Phanerozoic mobile belts of South-Western Europe. Despite extensive structural history studies, high-resolution seismic profiling has [...] Read more.
The Holy Cross Mountains represent an isolated outcrop of Palaeozoic rocks located in the Trans-European Suture Zone, which is the boundary between the Precambrian East European Craton and Phanerozoic mobile belts of South-Western Europe. Despite extensive structural history studies, high-resolution seismic profiling has not been applied to this region until now. This research introduces near-surface seismic imaging of the Holy Cross Fault, separating two tectonic units of different stratigraphic and deformation history. In our study, we utilize a carefully designed weight drop source survey with 5 m shot and receiver spacing and 4.5 Hz geophones. The imaging technique, combining seismic reflection profiling and travel time tomography, reveals detailed fault geometries down to 400 m. Precise data processing, including static corrections and noise attenuation, significantly enhanced signal-to-noise ratio and seismic resolution. Furthermore, the paper discusses various fault imaging techniques with their shortcomings. The data reveal a complex network of intersecting fault strands, confirming general thrust fault geometry of the fault system, that align with the region’s tectonic evolution. These findings enhance understanding of the Holy Cross Mountains’ structural framework and provide valuable reference data for future studies of similar tectonic environments. Full article
(This article belongs to the Special Issue Earthquake Engineering and Seismic Risk)
27 pages, 2285 KiB  
Article
Pinpointing Defects in Grounding Grids with Multistatic Radars
by Rodrigo M. S. de Oliveira and Pedro G. B. Maia
Energies 2025, 18(1), 150; https://doi.org/10.3390/en18010150 - 2 Jan 2025
Viewed by 214
Abstract
In this paper, we propose a method for locating discontinuities in grounding grids using a multistatic radar. The objective is to determine the fault position in the structure by injecting an ultra-wideband pulse (Gaussian monocycle) at one of the corners of the grid [...] Read more.
In this paper, we propose a method for locating discontinuities in grounding grids using a multistatic radar. The objective is to determine the fault position in the structure by injecting an ultra-wideband pulse (Gaussian monocycle) at one of the corners of the grid and analyzing the transient signals obtained at two sensors and at the transceiver. To perform the analysis and validation of the developed method, simulations based on the finite-difference time-domain (FDTD) technique were carried out to numerically solve Maxwell’s equations. The voltage signals obtained in an intact grounding grid are used as a reference. Differences between these reference voltages and the voltages obtained with the faulty grid are calculated. With these difference signals, the parameters of the radar ellipses and circle are obtained, which delimit the area where the fault can be found. These parameters depend on the wave propagation speed and the arrival times of the signals at the sensors and the transceiver. The results show that the proposed method is able to reduce the estimated fault location area to a range of 2% to 19% of the total grid area. In addition, the average distance between the actual fault and the center of the estimated region varies between 3.0 and 4.0 m. Full article
(This article belongs to the Special Issue Simulation and Analysis of Electrical Power Systems)
Show Figures

Figure 1

Figure 1
<p>The three-dimensional Yee cell used to discretize space and represent materials in the FDTD method.</p>
Full article ">Figure 2
<p>Ideal geometric configuration illustrating the operation of a multistatic radar. The lines of the circle and ellipses indicate the points where the target is.</p>
Full article ">Figure 3
<p>Parameters and reference points of an ellipse in the radar system.</p>
Full article ">Figure 4
<p>Grounding grid modeled in this work with the transceiver and receivers installed. The insets show details of the transceiver and receiver geometries. The grid is at a depth of 0.50 m.</p>
Full article ">Figure 5
<p>Gaussian monocycle function <math display="inline"><semantics> <mrow> <mi>G</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math> as a function of time <span class="html-italic">t</span>, describing the excitation source.</p>
Full article ">Figure 6
<p>Spectrum of the Gaussian monocycle function <math display="inline"><semantics> <mrow> <mi>G</mi> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>All faults studied in the grounding grid, whose actual positions are marked with the red symbol × and their respective fault indexes. The circles delimit the most probable regions for finding each respective fault, whose centers are the fault position estimates obtained through the multistatic radar method associated with the simplex method. In this case, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> in the soil.</p>
Full article ">Figure 8
<p>All faults studied in the grounding grid, whose actual positions are marked with the red symbol × and their respective fault indexes. The circles delimit the most probable regions for finding each respective fault, whose centers are the fault position estimates obtained through the multistatic radar method associated with the simplex method. In this case, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math> in the soil.</p>
Full article ">Figure 9
<p>All faults studied in the grounding grid, whose actual positions are marked with the red symbol × and their respective fault indexes. The circles delimit the most probable regions for finding each respective fault, whose centers are the fault position estimates obtained through the multistatic radar method associated with the simplex method. In this case, <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> in the soil.</p>
Full article ">Figure 10
<p>Graphs generated by the multistatic radar method applied to the grounding grid with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>. Discontinuities are identified as follows: (<b>a</b>) fault 1, (<b>b</b>) fault 2, (<b>c</b>) fault 3, (<b>d</b>) fault 4, (<b>e</b>) fault 5, (<b>f</b>) fault 6, (<b>g</b>) fault 7, (<b>h</b>) fault 8, and (<b>i</b>) fault 9.</p>
Full article ">Figure 11
<p>Graphs generated by the multistatic radar method applied to the grounding grid with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math>. Discontinuities are identified as follows: (<b>a</b>) fault 1, (<b>b</b>) fault 2, (<b>c</b>) fault 3, (<b>d</b>) fault 4, (<b>e</b>) fault 5, (<b>f</b>) fault 6, (<b>g</b>) fault 7, (<b>h</b>) fault 8, and (<b>i</b>) fault 9.</p>
Full article ">Figure 12
<p>Graphs generated by the multistatic radar method applied to the grounding grid with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>. Discontinuities are identified as follows: (<b>a</b>) fault 1, (<b>b</b>) fault 2, (<b>c</b>) fault 3, (<b>d</b>) fault 4, (<b>e</b>) fault 5, (<b>f</b>) fault 6, (<b>g</b>) fault 7, (<b>h</b>) fault 8, and (<b>i</b>) fault 9.</p>
Full article ">Figure 13
<p>Simulation results showing the magnitude of the electric field in the plane of the grounding grid for the time instant <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>1.33</mn> </mrow> </semantics></math> μs and <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>: (<b>a</b>) intact grid, (<b>b</b>) fault 1, (<b>c</b>) fault 2, and (<b>d</b>) fault 3. The discontinuities in sub-images (<b>b</b>–<b>d</b>) are highlighted within black squares.</p>
Full article ">Figure 14
<p>Voltages as a function of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>1</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 15
<p>Voltages as a function of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>2</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 16
<p>Voltages as a function of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>3</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 17
<p>Voltages as functions of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>1</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 18
<p>Voltages as functions of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>2</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 19
<p>Voltages as functions of time induced in receiver <math display="inline"><semantics> <msub> <mi>R</mi> <mn>3</mn> </msub> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>: (<b>a</b>) grid without faults, (<b>b</b>) grid with defect 01, and (<b>c</b>) difference between the voltage records obtained with and without defect 01.</p>
Full article ">Figure 20
<p>Fault locations by quadrant and path from the fault to the center of the corresponding FLEC for <math display="inline"><semantics> <mrow> <msub> <mi>ϵ</mi> <mi>r</mi> </msub> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math>. Exact fault locations are marked with the symbol × in red with respective fault indexes.</p>
Full article ">
18 pages, 9601 KiB  
Article
A Practical Approach for Fault Location in Transmission Lines with Series Compensation Using Artificial Neural Networks: Results with Field Data
by Simone Aparecida Rocha, Thiago Gomes de Mattos and Eduardo Gonzaga da Silveira
Energies 2025, 18(1), 145; https://doi.org/10.3390/en18010145 - 2 Jan 2025
Viewed by 304
Abstract
This paper presents a new method for fault location in transmission lines with series compensation, using data from voltage and current measurements at both terminals, applied to artificial neural networks. To determine the fault location, we present the proposal of using current phasors, [...] Read more.
This paper presents a new method for fault location in transmission lines with series compensation, using data from voltage and current measurements at both terminals, applied to artificial neural networks. To determine the fault location, we present the proposal of using current phasors, obtained from the oscillography recorded during the short circuit, as the input to the neural network for training. However, the method does not rely on the internal voltage values of the sources or their respective equivalent Thevenin impedances to generate training files for the neural network in a transient simulator. The source data are not known exactly at the time of the short circuit in the transmission line, leading to greater errors when neural networks are applied to real electrical systems of utility companies, which reduces the dependency on electrical network parameters. To present the new method, a conventional fault location algorithm based on neural networks is initially described, highlighting how the dependency on source parameters can hinder the application of the artificial neural network in real cases encountered in utility electrical systems. Subsequently, the new algorithm is described and applied to simulated and real fault cases. Low errors are obtained in both situations, demonstrating its effectiveness and practical applicability. It is noted that the neural networks used for real cases are trained using simulated faults but without any data from the terminal sources. Although we expect the findings of this paper to have relevance in transmission lines with series compensation, the new method can also be applied to conventional transmission lines, i.e., without series compensation, as evidenced by the results presented. Full article
(This article belongs to the Section F3: Power Electronics)
Show Figures

Figure 1

Figure 1
<p>Voltage–current characteristic of a MOV.</p>
Full article ">Figure 2
<p>Capacitor bank protection using varistors.</p>
Full article ">Figure 3
<p>MOV conduction during a short-circuit phase B.</p>
Full article ">Figure 4
<p>Flowchart of the fault location process.</p>
Full article ">Figure 5
<p>Compensated line transmission indicating Brazil’s electrical system.</p>
Full article ">Figure 6
<p>Voltages and currents for an AG fault. (<b>a</b>) Sobradinho Terminal; (<b>b</b>) S. J. Piauí Terminal.</p>
Full article ">Figure 7
<p>Currents in the transmission line, capacitor bank, and protective equipment for the AG fault.</p>
Full article ">Figure 8
<p>Elimination of source parameters.</p>
Full article ">Figure 9
<p>Algorithm proposed for fault location.</p>
Full article ">Figure 10
<p>Electrical system with an AG fault for the proposed algorithm.</p>
Full article ">Figure 11
<p>The selection of phasors to define the source values when assembling the circuit in the ATP.</p>
Full article ">Figure 12
<p>ATP model used for obtaining ANN training files.</p>
Full article ">Figure 13
<p>Models to simulated faults: (<b>a</b>) AG, (<b>b</b>) BC, (<b>c</b>) ABC, and (<b>d</b>) ACG (adapted from [<a href="#B1-energies-18-00145" class="html-bibr">1</a>]).</p>
Full article ">Figure 14
<p>Selection of input quantities in the ANN according to the type of fault.</p>
Full article ">Figure 15
<p>The selection of current phasors inputted into the ANN.</p>
Full article ">Figure 16
<p>Modular ANN structure for fault location.</p>
Full article ">Figure 17
<p>Voltages and currents for a BCG fault. (<b>a</b>) Sobradinho Terminal; (<b>b</b>) S. J. Piauí Terminal.</p>
Full article ">Figure 18
<p>Currents in the transmission line, capacitor bank, and protective equipment (BCG fault) (adapted from [<a href="#B1-energies-18-00145" class="html-bibr">1</a>]).</p>
Full article ">
16 pages, 4069 KiB  
Article
Photovoltaic Array Fault Diagnosis and Localization Method Based on Modulated Photocurrent and Machine Learning
by Yebo Tao, Tingting Yu and Jiayi Yang
Sensors 2025, 25(1), 136; https://doi.org/10.3390/s25010136 - 29 Dec 2024
Viewed by 345
Abstract
Photovoltaic arrays are exposed to outdoor conditions year-round, leading to degradation, cracks, open circuits, and other faults. Hence, the establishment of an effective fault diagnosis system for photovoltaic arrays is of paramount importance. However, existing fault diagnosis methods often trade off between high [...] Read more.
Photovoltaic arrays are exposed to outdoor conditions year-round, leading to degradation, cracks, open circuits, and other faults. Hence, the establishment of an effective fault diagnosis system for photovoltaic arrays is of paramount importance. However, existing fault diagnosis methods often trade off between high accuracy and localization. To address this concern, this paper proposes a fault identification and localization approach for photovoltaic arrays based on modulated photocurrent and machine learning. By irradiating different frequency-modulated light, this method separates photocurrent and directly measures the photoelectric conversion efficiency of each panel, achieving both high accuracy and localization. Through machine learning classification algorithms, the current amplitude and frequency of each photovoltaic panel are identified to achieve fault identification and localization. Compared to other methods, the strengths of this method lie in its ability to achieve high-speed and high-accuracy fault identification and localization by measuring only the short-circuit current. Additionally, the equipment cost is low. The feasibility of the proposed method is demonstrated through practical experimentation. It is determined that when utilizing a neural network algorithm, the fault identification speed meets measurement requirements (5800 obs/s), and the fault diagnosis accuracy is optimal (97.8%). Full article
(This article belongs to the Special Issue Fault Diagnosis for Photovoltaic Systems Based on Sensors)
Show Figures

Figure 1

Figure 1
<p>The photovoltaic array models. (<b>a</b>) Equivalent circuit of photovoltaic panel. (<b>b</b>) Schematic diagram of series photovoltaic array. (<b>c</b>) Schematic diagram of series-parallel photovoltaic array. (<b>d</b>) Schematic diagram of current spectrum generation.</p>
Full article ">Figure 2
<p>The simulation circuit and results of a series-connected photovoltaic array. (<b>a</b>) Series-connected simulation circuit. (<b>b</b>) Simulation results under the normal state. (<b>c</b>) Simulation results under degradation fault. (<b>d</b>) Simulation results under short-circuit fault. (<b>e</b>) Simulation results under open-circuit fault.</p>
Full article ">Figure 3
<p>The simulation circuit and results of a series-parallel connected photovoltaic array. (<b>a</b>) Series-parallel simulation circuit. (<b>b</b>) Simulation results under the normal state. (<b>c</b>) Simulation results under degradation fault. (<b>d</b>) Simulation results under short-circuit fault. (<b>e</b>) Simulation results under open-circuit fault.</p>
Full article ">Figure 4
<p>Experimental setup of modulated light generation and current detection. (<b>a</b>) Schematic diagram. (<b>b</b>) Block diagram. (<b>c</b>) Photograph.</p>
Full article ">Figure 5
<p>Photovoltaic array capable of switching among normal, open-circuit, short-circuit, and degradation states. (<b>a</b>) Circuit schematic diagram. (<b>b</b>) Photograph (front view). (<b>c</b>) Photograph (back view).</p>
Full article ">Figure 6
<p>Current spectrum under different states. (<b>a</b>) Normal state. (<b>b</b>) The photovoltaic panel is irradiated by modulated light at 10 kHz and is covered by half of the optical filter. (<b>c</b>) A photovoltaic panel is irradiated by modulated light at 10 kHz and is covered by one optical filter. (<b>d</b>) The photovoltaic panel is irradiated by modulated light at 10 kHz in a short-circuit state. (<b>e</b>) The photovoltaic panel is irradiated by modulated light at 10 kHz in an open-circuit state.</p>
Full article ">Figure 7
<p>Data set and results. (<b>a</b>) Distribution of the data set. (<b>b</b>) Neural Network architecture. (<b>c</b>) Testing confusion matrix (Neural Network). (<b>d</b>) Testing confusion matrix (Ensemble Learning: Bagging + Decision Tree).</p>
Full article ">
15 pages, 32385 KiB  
Technical Note
Aftershock Spatiotemporal Activity and Coseismic Slip Model of the 2022 Mw 6.7 Luding Earthquake: Fault Geometry Structures and Complex Rupture Characteristics
by Qibo Hu, Hongwei Liang, Hongyi Li, Xinjian Shan and Guohong Zhang
Remote Sens. 2025, 17(1), 70; https://doi.org/10.3390/rs17010070 - 28 Dec 2024
Viewed by 373
Abstract
On 5 September 2022, the moment magnitude (Mw) 6.7 Luding earthquake struck in the Xianshuihe Fault system on the eastern edge of the Tibet Plateau, illuminating the seismic gap in the Moxi segment. The fault system geometry and rupture process of this earthquake [...] Read more.
On 5 September 2022, the moment magnitude (Mw) 6.7 Luding earthquake struck in the Xianshuihe Fault system on the eastern edge of the Tibet Plateau, illuminating the seismic gap in the Moxi segment. The fault system geometry and rupture process of this earthquake are relatively complex. To better understand the underlying driving mechanisms, this study first uses the Interferometric Synthetic Aperture Radar (InSAR) technique to obtain static surface displacements, which are then combined with Global Positioning System (GPS) data to invert the coseismic slip distribution. A machine learning approach is applied to extract a high-quality aftershock catalog from the original seismic waveform data, enabling the analysis of the spatiotemporal characteristics of aftershock activity. The catalog is subsequently used for fault fitting to determine a reliable fault geometry. The coseismic slip is dominated by left-lateral strike-slip motion, distributed within a depth range of 0–15 km, with a maximum fault slip > 2 m. The relocated catalog contains 15,571 events. Aftershock activity is divided into four main seismic clusters, with two smaller clusters located to the north and south and four interval zones in between. The geometry of the five faults is fitted, revealing the complexity of the Xianshuihe Fault system. Additionally, the Luding earthquake did not fully rupture the Moxi segment. The unruptured areas to the north of the mainshock, as well as regions to the south near the Anninghe Fault, pose a potential seismic hazard. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Tectonic map of the Luding earthquake region. (<b>a</b>) The red star marks the mainshock epicenter in Luding, and the blue star marks the Kangding earthquake, while the green star indicates historical earthquakes with magnitudes ≥ 6.5 in the past 300 years. The colored triangles denote the durations recorded by broadband seismic stations during the study period. The violet squares represent cities, and the black dots indicate the distribution of detected seismic activity. (<b>b</b>) The blue arrow represents the horizontal displacement derived from the Global Positioning System (GPS). Blue dots indicate GPS stations. The inset map indicates the regional tectonics.</p>
Full article ">Figure 2
<p>Phase picking and association. (<b>a</b>) An example of earthquake association using USTC-Pickers to determine the arrival times of the P-wave (blue line) and S-wave (red line). Only the vertical component of the waveform is displayed, and it is band-pass filtered in the frequency range of 1–10 Hz; (<b>b</b>) travel time–hypocentral distance curves for the associated events. The black line represents the fitted approximate velocities of the P-wave (6.0 km/s) and S-wave (3.5 km/s).</p>
Full article ">Figure 3
<p>Comparison between the detected catalog and the manual catalog and the distribution of seismic release energy. (<b>a</b>) Distribution of magnitude differences between coexisting events after magnitude correction for both catalogs. (<b>b</b>) Characteristics of the magnitude–frequency distribution for the manual and detected catalogs. The red bars represent events identified by machine learning, while the blue bars correspond to events identified manually. (<b>c</b>) Distribution of seismic energy, with a grid size of 0.01° × 0.01°. The red star marks the location of the mainshock.</p>
Full article ">Figure 4
<p>Sentinel-1 interferograms and line-of-sight (LOS) displacements from descending track. (<b>a</b>,<b>b</b>) The red stars indicate the location of the Luding earthquake.</p>
Full article ">Figure 5
<p>Spatial distribution of aftershocks at different times. (<b>a</b>–<b>e</b>) Aftershock activity is shown by colored dots, with the color representing depth. The red star is the location of the mainshock. The black circles indicate divided aftershock groups (S1–S4), and the black lines divide aftershock gaps (G1–G4). (<b>f</b>) Spatial–temporal aftershock evolution. The solid black line shows the direction of expansion of seismic activity over time.</p>
Full article ">Figure 6
<p>Regional fault profile. (<b>a</b>) The dashed blue line indicates the cross-section profiles, the red star indicates the location of the mainshock, and the black triangle indicates the Gongga Mountain. The red surface indicates the fault plane used to invert the coseismic slip. (<b>b</b>–<b>k</b>) The solid black line indicates the fault fit, and the parameters are given in the text. The dashed black line indicates the trend of changes in the depth of seismic activity. The time in hours until aftershocks occur is indicated by the color.</p>
Full article ">Figure 7
<p>Coseismic slip model and shear stress changes for the Luding earthquake. (<b>a</b>) The blue arrows indicate the direction of the slip, with the length representing the amount of slippage. The gray dashed lines represent contours of equal slippage. (<b>b</b>) The red star indicates the epicenter, and the red beach ball indicates the source mechanism solution.</p>
Full article ">Figure 8
<p>The relationship between coseismic slip and the distribution of aftershocks at different times. (<b>a</b>–<b>f</b>) The steel-blue dots indicate aftershocks; The gray dashed lines represent contours of equal slippage;Red star indicates the location of the Luding earthquake.</p>
Full article ">Figure 9
<p>A 3D resistivity structure depth slice. (<b>a</b>) Black dots indicate aftershocks, and red stars indicate the mainshock. The white arrow indicates the stress direction derived from the inversion of the source mechanism. (<b>b</b>–<b>f</b>) Cross-sections of resistivity structure.</p>
Full article ">
16 pages, 5118 KiB  
Article
Prediction of GNSS Velocity Accuracies Using Machine Learning Algorithms for Active Fault Slip Rate Determination and Earthquake Hazard Assessment
by Halil İbrahim Solak
Appl. Sci. 2025, 15(1), 113; https://doi.org/10.3390/app15010113 - 27 Dec 2024
Viewed by 364
Abstract
GNSS technology utilizes satellite signals to determine the position of a point on Earth. Using this location information, the GNSS velocities of the points can be calculated. GNSS velocity accuracies are crucial for studies requiring high precision, as fault slip rates typically range [...] Read more.
GNSS technology utilizes satellite signals to determine the position of a point on Earth. Using this location information, the GNSS velocities of the points can be calculated. GNSS velocity accuracies are crucial for studies requiring high precision, as fault slip rates typically range within a few millimeters per year. This study employs machine learning (ML) algorithms to predict GNSS velocity accuracies for fault slip rate estimation and earthquake hazard analysis. GNSS data from four CORS stations collected over 1-, 2-, and 3-year intervals with observation durations of 2, 4, 6, 8, and 12 h, were analyzed to generate velocity estimates. Position accuracies, observation intervals, and corresponding velocity accuracies formed two datasets for the East and North components. ML models, including Support Vector Machine, Random Forest, K-Nearest Neighbors, and Multiple Linear Regression, were used to model the relationship between position and velocity accuracies. The findings reveal that the Random Forest, which makes more accurate and reliable predictions by evaluating many decision trees together, achieved over 90% accuracy for both components. Velocity accuracies of ±1.3 mm/year were obtained for 1-year interval data, while accuracies of ±0.6 mm/year were achieved for the 2- and 3-year intervals. Three campaigns were deemed sufficient for Holocene faults with higher slip rates. However, for Quaternary faults with lower slip rates, longer observation periods or additional campaigns are necessary to ensure reliable velocity estimates. This highlights the need for GNSS observation planning based on fault activity. Full article
(This article belongs to the Section Earth Sciences)
Show Figures

Figure 1

Figure 1
<p>Locations of the stations.</p>
Full article ">Figure 2
<p>Time series of ESKS station.</p>
Full article ">Figure 3
<p>Number of neighbors determined for KNN according to elbow method.</p>
Full article ">Figure 4
<p>Reference velocity accuracies and the residuals in RF for test dataset (The red line represents the referenced GNSS velocity accuracies, while the blue line represents the residuals).</p>
Full article ">Figure 5
<p>Reference velocity accuracies and the residuals in KNN for test dataset (The red line represents the referenced GNSS velocity accuracies, while the blue line represents the residuals).</p>
Full article ">Figure 6
<p>Reference velocity accuracies and the residuals in SVM for test dataset (The red line represents the referenced GNSS velocity accuracies, while the blue line represents the residuals).</p>
Full article ">Figure 7
<p>Reference velocity accuracies and the residuals in MLR for test dataset (The red line represents the referenced GNSS velocity accuracies, while the blue line represents the residuals).</p>
Full article ">Figure 8
<p>Visualization of the test results for the developed models on campaign-based external GNSS datasets.</p>
Full article ">Figure 9
<p>Prediction of velocity accuracies as a function of position accuracies and temporal intervals.</p>
Full article ">
22 pages, 6291 KiB  
Article
Origin of the Miaoling Gold Deposit, Xiong’ershan District, China: Findings Based on the Trace Element Characteristics and Sulfur Isotope Compositions of Pyrite
by Simo Chen, Junqiang Xu, Yanchen Yang, Shijiong Han, Peichao Ding, Zhaoyang Song, Tianwen Chen and Daixin Zhang
Minerals 2025, 15(1), 6; https://doi.org/10.3390/min15010006 - 24 Dec 2024
Viewed by 300
Abstract
The Xiong’ershan district is situated on the southern margin of the North China Craton (NCC) and located within the Qinling–Dabieshan Orogen’s orogenic zone. It is adjacent to the XiaoQinling mining district and exhibits very favorable geological conditions for mineralization, as the district contains [...] Read more.
The Xiong’ershan district is situated on the southern margin of the North China Craton (NCC) and located within the Qinling–Dabieshan Orogen’s orogenic zone. It is adjacent to the XiaoQinling mining district and exhibits very favorable geological conditions for mineralization, as the district contains numerous gold deposits, positioning it as one of the key gold-producing areas of China. The Miaoling gold deposit is a hydrothermal deposit and is controlled by the Mesozoic nearly NS-trending fault. The ore bodies are hosted in the Mesoproterozoic Xiong’er Group of the Changcheng System of volcanic rocks, with reserves reaching large-scale levels. Pyrite is the main gold-bearing mineral and can be classified into four generations: early-stage fine- to medium-grained euhedral to subhedral cubic pyrite (Py1); medium- to coarse-grained euhedral to subhedral cubic granular pyrite in quartz veins (Py2a); fine-grained subhedral to anhedral disseminated pyrite in altered rocks (Py2b); and late-stage anhedral granular and fine-veinlet pyrite in later quartz veins (Py3). Through in situ trace element analysis of the pyrite using LA-ICP-MS, a positive correlation between Au and As was observed during the main mineralization stage; gold mainly exists as a solid solution within the pyrite lattice, and the ablation signal curve reflecting the intensity of trace element signals showed that gold also occurs as micron-scale mineral inclusions. The trace element content suggested a gradual increase in oxygen fugacity from Stage 1 to Stage 2, followed by a decrease from Stage 2 to Stage 3. The Co/Ni values in the pyrite (0.56 to 62.02, with an average of 12.34) exhibited characteristics of magmatic hydrothermal pyrite. The in situ sulfur isotope analysis of the pyrite using LA-MC-ICP-MS showed δ34S values of 4.24‰ for Stage 1, −6.63‰ to −13.79‰ for Stage 2, and −4.31‰ to −5.15‰ for Stage 3. Considering sulfur isotope fractionation, the δ34S value of the hydrothermal fluid during the main mineralization stage was calculated to be between 0.31‰ and 2.68‰. Full article
(This article belongs to the Special Issue The Formation and Evolution of Gold Deposits in China)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Map of the distribution and locations of plates in China [<a href="#B25-minerals-15-00006" class="html-bibr">25</a>]. (<b>b</b>) Geological map of the southern margin of the NCC [<a href="#B5-minerals-15-00006" class="html-bibr">5</a>]. (<b>c</b>) Regional geological map of the Xiong’ershan district [<a href="#B7-minerals-15-00006" class="html-bibr">7</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Geological diagram of Miaoling gold deposit [<a href="#B12-minerals-15-00006" class="html-bibr">12</a>]. (<b>b</b>) Projection of ore body in Miaoling gold deposit [<a href="#B39-minerals-15-00006" class="html-bibr">39</a>]. (<b>c</b>) Exploration line profile of Miaoling gold deposit [<a href="#B40-minerals-15-00006" class="html-bibr">40</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) F8 fault structure alteration zone. (<b>b</b>,<b>c</b>) Primary ore body in the F8 fault.</p>
Full article ">Figure 4
<p>Field and underground photos of quartz veins and altered rocks in four stages of the Miaoling gold deposit: (<b>a</b>) Stage 1 quartz vein; (<b>b</b>) two stages of crossing–cutting quartz veins; (<b>c</b>) Stage 2 quartz vein and pyrite; (<b>d</b>) pyrite in altered rocks; (<b>e</b>) Stage 3 quartz–polymetallic sulfide vein; (<b>f</b>) specularite vein; (<b>g</b>) Stage 4 quartz–fluorite vein; (<b>h</b>) carbonate vein; (<b>i</b>) epidotization; Py—pyrite; Py2a—pyrite in quartz vein; Py2b—pyrite in altered rocks; Ccp—chalcopyrite; Sp—sphalerite; Gn—galena; Spe—specularite; Qtz—quartz; Fl—fluorite; Cal—calcite; Ep—epidotization.</p>
Full article ">Figure 5
<p>An overview of the paragenetic sequence from the four stages of mineralization in the Miaoling gold deposit.</p>
Full article ">Figure 6
<p>Microscopic photos of pyrite in Miaoling gold deposit: (<b>a</b>,<b>b</b>) pyrite in Stage 1 quartz vein; (<b>c</b>–<b>d</b>) pyrite in Stage 2 quartz vein; (<b>e</b>–<b>f</b>) pyrite in altered rocks; (g) pyrite in Stage 3 quartz vein; (<b>h</b>–<b>i</b>) polymetallic sulfide in Stage 3 quartz vein. Py—pyrite; Gn—galena; Ccp—chalcopyrite; Sp—sphalerite.</p>
Full article ">Figure 7
<p>Box and whisker plots of trace element concentrations in pyrite, acquired by means of LA-ICP-MS analysis, from the Miaoling gold deposit.</p>
Full article ">Figure 8
<p></p>
Full article ">Figure 9
<p></p>
Full article ">Figure 10
<p>(<b>a</b>) A histogram of the sulfur isotopic compositions of sulfides from the Miaoling gold deposit. (<b>b</b>) The ranges of δ<sup>34</sup>S values in the Miaoling gold deposit and other deposits [<a href="#B5-minerals-15-00006" class="html-bibr">5</a>,<a href="#B12-minerals-15-00006" class="html-bibr">12</a>,<a href="#B84-minerals-15-00006" class="html-bibr">84</a>,<a href="#B87-minerals-15-00006" class="html-bibr">87</a>,<a href="#B88-minerals-15-00006" class="html-bibr">88</a>,<a href="#B91-minerals-15-00006" class="html-bibr">91</a>].</p>
Full article ">
16 pages, 14248 KiB  
Article
Holocene Activity Characteristics and Seismic Risk of Major Earthquakes in the Middle Segment of the Jinshajiang Fault Zone, East of the Qinghai–Tibetan Plateau
by Mingjian Liang, Naifei Luo, Yunxi Dong, Ling Tan, Jinrong Su and Weiwei Wu
Appl. Sci. 2025, 15(1), 9; https://doi.org/10.3390/app15010009 - 24 Dec 2024
Viewed by 284
Abstract
The Jinshajiang fault zone is the western boundary fault of the Sichuan–Yunnan block, located east of the Qinghai–Tibetan Plateau. It is a complex tectonic suture belt with multi-phase activity and is characterized by multiple sets of parallel or intersecting faults. Using high-resolution image [...] Read more.
The Jinshajiang fault zone is the western boundary fault of the Sichuan–Yunnan block, located east of the Qinghai–Tibetan Plateau. It is a complex tectonic suture belt with multi-phase activity and is characterized by multiple sets of parallel or intersecting faults. Using high-resolution image interpretation, seismic geological surveys, and trench studies, we examined the Holocene activity and obtained the paleoseismic sequences on the middle segment of the fault zone. Thus, we could analyze the kinematic characteristics of the fault and its potential risk of strong earthquakes. Our results indicated that the predominant movement of the fault zone was strike-slip motion. In the Jinshajiang fault zone, the Late Quaternary horizontal slip rates of the north-northeast-trending Yarigong fault and the northeast-trending Ciwu fault were 3.6 ± 0.6 mm/a and 2.5 ± 0.5 mm/a, respectively. Three paleoseismic events were identified on the Yarigong fault, dated 6745–3848, 3742–1899, and 1494–1112 cal BP, and on the Ciwu fault, constrained to 32,566–29,430, 24,056–22,990, and 2875–2723 cal BP. The last major earthquake on the Ciwu fault occurred approximately 2800 years ago; therefore, its future seismic hazard deserves attention. Full article
(This article belongs to the Special Issue Paleoseismology and Disaster Prevention)
Show Figures

Figure 1

Figure 1
<p>Tectonic framework and seismic activity of the Jinshajiang fault zone and adjacent regions. The black dashed rectangle represents the study area.</p>
Full article ">Figure 2
<p>(<b>a</b>) Distribution map of major faults and field survey sites in the middle section of the Jinshajiang fault zone and its adjacent areas. (<b>b</b>) On the basis of the GF-7 satellite data, a hillshade map was generated to interpret the detailed fault tracks of the Yarigong and Ciwu faults.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Geological profile of the Jinshajiang fault zone in Hongdong Village. (<b>c</b>) Photograph of the top of the geological section showing the fault displaced top strata.</p>
Full article ">Figure 4
<p>Right-lateral offset of a gully and profiles of the dating sample collection south of Yarigong Town. (<b>a</b>,<b>b</b>) The right-lateral offset of the gully was obtained using unmanned aerial vehicle (UAV) photogrammetry. The white rectangle in panel (<b>b</b>) indicates the sampling location of the dating samples. (<b>c</b>,<b>d</b>) The dating samples were collected from the T3 and T4 terraces of the Muqu River, respectively.</p>
Full article ">Figure 5
<p>Fault profiles exposed along the Yarigong fault. (<b>a</b>–<b>c</b>) The fault profiles are exposed north of the Ran, Dalong, and Lide villages, respectively, where the fault has displaced Late Quaternary strata. In (<b>a</b>), the fault has displaced the Holocene alluvial layer. And the fault has displaced the late Pleistocene alluvial layer in (<b>b</b>). The red arrows indicate the fault traces.</p>
Full article ">Figure 6
<p>Photograph of the fault profile and its explanatory profile.</p>
Full article ">Figure 7
<p>Structural landform of the Ciwu fault southwest of Nidou Village. (<b>a</b>) Fault track of the newest activity of the Ciwu fault on the alluvial fan. (<b>b</b>,<b>c</b>) Linear fault trough gullies and reverse fault scarps. The red arrows indicate the fault traces.</p>
Full article ">Figure 8
<p>(<b>a</b>) Right-lateral offset of the T2 terrace of the Ciwu River and location of the dating sample. The image was obtained using UAV photogrammetry. (<b>b</b>) Photograph showing the dating sample collection section in the T2 terrace of the Ciwu River.</p>
Full article ">Figure 9
<p>(<b>a</b>) Tectonic landform near trench TC1 according to UAV photogrammetry. (<b>b</b>) Photograph of the linear fault trough landform. The red arrows indicate the fault trace.</p>
Full article ">Figure 10
<p>Photograph of the southern wall of the Bugge trench and explanatory profile. The radiocarbon dating sample ages of the TC1 trench are detailed in <a href="#applsci-15-00009-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 11
<p>(<b>a</b>) Tectonic landform near trench TC2 shown via UAV photogrammetry. The gully to the south of the site is right-laterally offset by approximately 18 ± 2 m. (<b>b</b>) Photograph of the trench site.</p>
Full article ">Figure 12
<p>Photograph of the southern wall of the Bugge trench and explanatory profile. The radiocarbon dating sample ages of the TC2 trench are detailed in <a href="#applsci-15-00009-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 13
<p>Paleoseismic sequences of the Yarigong and Ciwu faults, the raw carbon ages of which were calibrated using OxCal 4.4.4 [<a href="#B42-applsci-15-00009" class="html-bibr">42</a>]. (<b>a</b>,<b>b</b>) are the paleoseismic sequences revealed in trench TC1 and TC2, respectively.</p>
Full article ">
19 pages, 32077 KiB  
Article
Present-Day Tectonic Deformation Characteristics of the Northeastern Pamir Margin Constrained by InSAR and GPS Observations
by Junjie Zhang, Xiaogang Song, Donglin Wu and Xinjian Shan
Remote Sens. 2024, 16(24), 4771; https://doi.org/10.3390/rs16244771 - 21 Dec 2024
Viewed by 387
Abstract
The Pamir is located on the northwestern margin of the Tibetan Plateau, which is an area of intense continental deformation and part of the famous India–Himalaya collision zone. The dominant structural deformation in the eastern Pamir is characterized by a 250 km long [...] Read more.
The Pamir is located on the northwestern margin of the Tibetan Plateau, which is an area of intense continental deformation and part of the famous India–Himalaya collision zone. The dominant structural deformation in the eastern Pamir is characterized by a 250 km long east–west extensional fault system, known as the Kongur Shan extensional system (KSES), which has developed a series of faults with different orientations and characteristics, resulting in highly complex structural deformation and lacking sufficient geodetic constraints. We collected Sentinel-1 SAR data from December 2016 to March 2023, obtained high-resolution ascending and descending LOS velocities and 3D deformation fields, and combined them with GPS data to constrain the current motion characteristics of the northeastern Pamirs for the first time. Based on the two-dimensional screw dislocation model and using the Bayesian Markov chain Monte Carlo (MCMC) inversion method, the kinematic parameters of the fault were calculated, revealing the fault kinematic characteristics in this region. Our results demonstrate that the present-day deformation of the KSES is dominated by nearly E–W extension, with maximum extensional motion concentrated in its central segment, reaching peak extension rates of ~7.59 mm/yr corresponding to the Kongur Shan. The right-lateral Muji fault at the northern end exhibits equivalent rates of extensional motion with a relatively shallow locking depth. The strike-slip rate along the Muji fault gradually increases from west to east, ranging approximately between 4 and 6 mm/yr, significantly influenced by the eastern normal fault. The Tahman fault (TKF) at the southernmost end of the KSES shows an extension rate of ~1.5 mm/yr accompanied by minor strike-slip motion. The Kashi anticline is approaching stability, while the Mushi anticline along the eastern Pamir frontal thrust (PFT) remains active with continuous uplift at ~2 mm/yr, indicating that deformation along the Tarim Basin–Tian Shan boundary has propagated southward from the South Tian Shan thrust (STST). Overall, this study demonstrates the effectiveness of integrated InSAR and GPS data in constraining contemporary deformation patterns along the northeastern Pamir margin, contributing to our understanding of the region’s tectonic characteristics. Full article
Show Figures

Figure 1

Figure 1
<p>Tectonics and seismicity of the study area. (<b>a</b>) The yellow and red rectangle shows the spatial footprint of the Sentinel-1 InSAR coverage. Blue arrows show the GPS horizontal velocity field with respect to the stable Eurasian plate [<a href="#B28-remotesensing-16-04771" class="html-bibr">28</a>]. Circles of different colors represent earthquake events of varying magnitudes. (<b>b</b>) Fault structures in the eastern part of PFT. Red and pink focal mechanisms represent the mainshock and aftershocks of the 1985 Wuqia earthquake. Brown focal mechanisms represent the mainshock of the 2016 Aketao earthquake. (<b>c</b>) Fault segmentation in KESE. S1–S5 correspond to different segments, respectively. STST = southern Tian Shan thrust, PFT = Pamir frontal thrust, KSES = Kongur Shan extensional system, MPT = main Pamir thrust, KKF = Karakax fault, KYTS = Kashgar–Yecheng transfer system, TFF = Talas–Fergana fault, TT = Tuomuluoan thrust, MF = Muji fault, KATF = King Ata Tagh normal fault, KSF = Kongur Shan normal fault, MAF = Muztagh Ata normal fault, TF = Tahman normal fault, TKF = Tashkorgan normal fault.</p>
Full article ">Figure 2
<p>The Sentinel-1 A/B data processing workflow. It consists of three steps, including interferograms generation, SBAS time series analysis, and three-dimensional deformation field solution.</p>
Full article ">Figure 3
<p>Perpendicular and temporal baseline plot showing the network of interferograms on one ascending track (<b>a</b>) and one descending track (<b>b</b>) used in this study. The number of total interferograms are labelled for each track.</p>
Full article ">Figure 4
<p>Interpolated GNSS velocities map. The interpolated GNSS velocities using the method outlined by Shen et al. [<a href="#B47-remotesensing-16-04771" class="html-bibr">47</a>], (<b>a</b>) corresponds to EW and (<b>b</b>) corresponds to NS. GNSS velocities are resampled to a resolution of 0.01 degrees. Different colored circles represent different GPS data, and the color bars for GPS various data and the interpolated velocity field are identical.</p>
Full article ">Figure 5
<p>The satellite line-of-sight (LOS) velocity fields of the northeastern Pamir margin. Red lines represent the fault crossing profiles, each profile for 60 km long and 10 km wide, distributed along six sub-faults of the KSES. (<b>a</b>) corresponds to ascending track and (<b>b</b>) corresponds to descending track.</p>
Full article ">Figure 6
<p>Joint InSAR-GPS three-dimensional deformation field. (<b>a</b>–<b>c</b>) are east–west, north–south, and vertical components, respectively.</p>
Full article ">Figure 7
<p>GPS profiles and results. (<b>a</b>) Four profiles in KSES. Each profile is 300 km long and 50 km wide. (<b>b</b>) The GPS data was projected parallel to and perpendicular to the local fault, respectively, and combined with fault strike.</p>
Full article ">Figure 8
<p>The cross-fault profiles of ascending and descending LOS deformation velocities. (<b>a</b>) represents ascending LOS velocity profiles, and (<b>b</b>) represents descending LOS velocity profiles. (<b>a</b>–<b>f</b>) correspond successively to the six profiles in <a href="#remotesensing-16-04771-f005" class="html-fig">Figure 5</a>. Black dots are binned average values every 1 km along the profile. Gray vertical stripes indicate the mountains on profiles. Red and purple lines are the best-fitting models. The black dotted line indicates the fault location.</p>
Full article ">Figure 9
<p>An example of Bayesian MCMC inversion results for profile aa’. Posterior marginal probability density functions illustrating parameter estimation and uncertainty quantification. (<b>Top</b>): profile aa’ topography from the Copernicus DEM data with 30 m spatial resolution (average elevation: white line; min/max: gray lines). (<b>Middle</b>): InSAR LOS velocities with the best-fitting predicted velocities. (<b>Bottom</b>): model-predicted fault-parallel, fault-normal, and vertical velocities.</p>
Full article ">Figure 10
<p>The LOS velocities and vertical component of the eastern PFT. (<b>a</b>–<b>c</b>) are ascending, descending, and vertical components, respectively. The abnormal deformation area corresponding to the black circle and black rectangle are caused by industrial activity. The satellite images corresponding to these two regions are shown in <a href="#app1-remotesensing-16-04771" class="html-app">Figure S5</a>. (<b>d</b>–<b>f</b>) correspond to the results of profiles aa’, bb’, and cc’ in (<b>a</b>–<b>c</b>), respectively. The pink, blue, and green points correspond to the results of profiles aa’, bb’, and cc’, respectively.</p>
Full article ">
19 pages, 5843 KiB  
Article
Identification of Strike-Slip Faults and Their Control on the Permian Maokou Gas Reservoir in the Southern Sichuan Basin (SW China): Fault Intersections as Hydrocarbon Enrichment Zones
by Jiawei Liu, Guanghui Wu, Hai Li, Wenjin Zhang, Majia Zheng, Hui Long, Chenghai Li and Min Deng
Energies 2024, 17(24), 6438; https://doi.org/10.3390/en17246438 - 20 Dec 2024
Viewed by 350
Abstract
The Middle Permian Maokou Formation carbonate rocks in the southern Sichuan Basin are import targets for hydrocarbon exploration, with numerous gas fields discovered in structural traps. However, as exploration extends into slope and syncline zones, the limestone reservoirs become denser, and fluid distribution [...] Read more.
The Middle Permian Maokou Formation carbonate rocks in the southern Sichuan Basin are import targets for hydrocarbon exploration, with numerous gas fields discovered in structural traps. However, as exploration extends into slope and syncline zones, the limestone reservoirs become denser, and fluid distribution becomes increasingly complex, limiting efficient exploration and development. Identifying the key factors controlling natural gas accumulation is therefore critical. This study is the first to apply deep learning techniques to fault detection in the southern Sichuan Basin, identifying previously undetected WE-trending subtle strike-slip faults (vertical displacement < 20 m). By integrating well logging, seismic, and production data, we highlight the primary factors influencing natural gas accumulation in the Maokou Formation. The results demonstrate that 80% of production comes from less than 30% of the well, and that high-yield wells are strongly associated with faults, particularly in slope and syncline zones where such wells are located within 200 m of fault zones. The faults can increase the drilling leakage of the Maokou wells by (7–10) times, raise the reservoir thickness to 30 m, and more than double the production. Furthermore, 73% of high-yield wells are concentrated in areas of fault intersection with high vertical continuity. Based on these insights, we propose four hydrocarbon enrichment models for anticline and syncline zones. Key factors controlling gas accumulation and high production include fault intersections, high vertical fault continuity, and local structural highs. This research demonstrates the effectiveness of deep learning for fault detection in complex geological settings and enhances our understanding of fault systems and carbonate gas reservoir exploration. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Tectonic divisions of the Sichuan Basin in China and location of the study area; (<b>b</b>) Surface geological map of the southern Sichuan Basin; (<b>c</b>) Geological section across southern Sichuan Basin; (<b>d</b>) Tectonic-stratigraphic column in the southern Sichuan Basin (after reference [<a href="#B19-energies-17-06438" class="html-bibr">19</a>]).</p>
Full article ">Figure 2
<p>(<b>a</b>) The seismic section showing the impact of evaporate and fault scale on the imaging of the deep fault; (<b>b</b>) the planar coherence at the bottom of the Cambrian showing only NNE-trending thrust faults in the Luzhou area.</p>
Full article ">Figure 3
<p>The architecture of the Unet (modified from [<a href="#B34-energies-17-06438" class="html-bibr">34</a>]).</p>
Full article ">Figure 4
<p>The seismic section of coherence (<b>a</b>,<b>c</b>) and DL attribute (<b>b</b>,<b>d</b>) showing the fault. (The section location is displayed in <a href="#energies-17-06438-f005" class="html-fig">Figure 5</a>.)</p>
Full article ">Figure 5
<p>Planar coherence attributes of coherence and DL attribute at the bottom of the Cambrian (<b>a</b>,<b>b</b>), and the bottom of the Upper Permian (<b>c,d</b>). (The location is displayed in <a href="#energies-17-06438-f006" class="html-fig">Figure 6</a>.)</p>
Full article ">Figure 6
<p>The interpretation sketch of fault system at (<b>a</b>) the bottom of the Upper Permian, (<b>b</b>) the bottom of the Upper Ordovician and (<b>c</b>) the bottom of the Cambrian in the Luzhou area (Є<sub>1</sub>q: base of the Cambrian; O<sub>3</sub>w: base of the Upper Ordovician; P<sub>2</sub>l: base of the Upper Permian).</p>
Full article ">Figure 7
<p>The thrust system in the Luzhou area (Є<sub>2</sub>g: base of the middle Cambrian; O<sub>3</sub>w: base of the Upper Ordovician; P<sub>1</sub>l: base of the Permian; P<sub>2</sub>l: base of the Upper Permian; T<sub>1</sub>f: base of the Lower Triassic; T<sub>1</sub>j: base of the Lower Triassic Jialingjing Formation; T<sub>3</sub>x: base of the Upper Triassic; The section location is displayed in <a href="#energies-17-06438-f006" class="html-fig">Figure 6</a>).</p>
Full article ">Figure 8
<p>The WE-trending (<b>a</b>,<b>b</b>) and NW-trending (<b>c</b>) fault system in the Luzhou area (O<sub>3</sub>w: base of the Upper Ordovician; P<sub>1</sub>l: base of the Permian; P<sub>2</sub>l: base of the Upper Permian; The section location is displayed in <a href="#energies-17-06438-f006" class="html-fig">Figure 6</a>).</p>
Full article ">Figure 9
<p>The cumulative gas and water production of the Maokou Formation in the Luzhou area. (High-yield well: The cumulative gas production &gt; 1 × 10<sup>8</sup> m<sup>3</sup>; Medium-yield well: 0.5 × 10<sup>8</sup> m<sup>3</sup> &lt; The cumulative gas production &lt; 1 × 10<sup>8</sup> m<sup>3</sup>; Low-yield well: The cumulative gas production &lt; 0.5 × 10<sup>8</sup> m<sup>3</sup>).</p>
Full article ">Figure 10
<p>Production curve of high-yield well of Maokou reservoir in the Luzhou area. (<b>a</b>) Gas production of Well Y71 in anticline zone; (<b>b</b>) Gas production of Well J41 in slope zone; (<b>c</b>) Gas production of Well J002-x2 in syncline zone.</p>
Full article ">Figure 11
<p>Hydrocarbon accumulation model of fault-controlled gas system in the southern Sichuan Basin. (Strike-slip faults serve as favorable lateral and vertical pathways for oil and gas migration and accumulation in the syncline zones, while thrust faults are advantageous channels for oil and gas migration in the anticline zones. Additionally, local traps can be separated by strike-slip faults.)</p>
Full article ">Figure 12
<p>DL attribute displays the vertical continuity structure of thrust (<b>a</b>–<b>c</b>) and strike-slip fault (<b>d</b>–<b>f</b>). From left to right, the continuity of fault gradually increases.</p>
Full article ">Figure 13
<p>(<b>a</b>) Relationship between cumulative gas production and distance to fault; (<b>b</b>) Relationship between fault direction and cumulative gas production; (<b>c</b>) Relationship between reservoir thickness and distance to fault; (<b>d</b>) Relationship between drilling leakage and distance to fault.</p>
Full article ">Figure 14
<p>The enrichment model of Maokou gas reservoir controlled by fault in the Luzhou area. (<b>a</b>) Enrichment model of strike-slip faults and thrust faults intersecting in anticline zone; (<b>b</b>) Enrichment model of thrust faults and the Permian reverse faults intersecting in anticline zone; (<b>c</b>) Enrichment model of strike-slip faults uplift in syncline zone; (<b>d</b>) Enrichment model of the Permian reverse faults and strike-slip faults intersecting in syncline zone.</p>
Full article ">
23 pages, 5302 KiB  
Article
A Novel Method for Automatically and Accurately Diagnosing Demagnetization Fault in Direct-Drive PMSMs Using Three PNNs
by Yiyong Xiong, Jinghong Zhao, Sinian Yan, Kun Wei and Haiwen Zhou
Appl. Sci. 2024, 14(24), 11943; https://doi.org/10.3390/app142411943 - 20 Dec 2024
Viewed by 388
Abstract
Direct-drive permanent magnet synchronous machines (DDPMSMs) have recently become an ideal candidate for applications such as military, robotics, electric vehicles, etc. These machines eliminate the need for a transmission mechanism and excitation coil circuits, which enhances the system’s overall efficiency and decreases the [...] Read more.
Direct-drive permanent magnet synchronous machines (DDPMSMs) have recently become an ideal candidate for applications such as military, robotics, electric vehicles, etc. These machines eliminate the need for a transmission mechanism and excitation coil circuits, which enhances the system’s overall efficiency and decreases the likelihood of failures. However, it may incur demagnetization faults. Due to the characteristic of having a large number of pole pairs, this type of machine exhibits numerous demagnetization fault modes, which poses a challenge in locating demagnetization faults. This paper proposed a probabilistic neural network (PNN)-based diagnostic system to detect and locate demagnetization faults in DDPMSMs, using information obtained through three toroidal-yoke-type search coils arranged at the bottom of the stator slot. A rotor partition method is proposed to solve the problem of demagnetization fault location difficulty caused by various fault modes. Demagnetization fault location is achieved by sequentially diagnosing the condition of each partition of permanent magnets. Three demagnetization fault identified signals (DFISs) are constructed by the voltage of the three toroidal-yoke coils, which are used as inputs of PNNs. Three PNNs have been designed to map the extracted features and their corresponding types of demagnetization faults. The database for training and testing the PNNs is generated from a DDPMSM with different demagnetization conditions and different operating conditions, which are established through an experimentally validated mathematical model, an FEM model, and experiments. The simulation and experimental test results showed that the accuracy in diagnosing the location of the demagnetization fault is 99.2% when the demagnetization severity is 10%, which demonstrates the effectiveness of the proposed three PNN-based diagnostic approach. Full article
Show Figures

Figure 1

Figure 1
<p>Placement scheme of toroidal-yoke-search coil.</p>
Full article ">Figure 2
<p>The waveforms of <span class="html-italic">SI</span><sub>m−2</sub>. (<b>a</b>) Mode 1 to 6. (<b>b</b>) Modes 6, 7, and 8.</p>
Full article ">Figure 3
<p>The waveforms of <span class="html-italic">SI</span><sub>m−2</sub> under Mode 7 with 100% demagnetization and Mode 8 with 50% demagnetization.</p>
Full article ">Figure 4
<p>Normalized <span class="html-italic">SI</span><sub>m</sub> under six different demagnetization fault modes.</p>
Full article ">Figure 5
<p>The waveforms of normalized <span class="html-italic">SI</span><sub>a1-2</sub>. (<b>a</b>) Mode 6H, Mode 7H, Mode 6D, Mode 8H, and Mode 8D. (<b>b</b>) Mode 7D and Mode 8H.</p>
Full article ">Figure 6
<p>The waveforms of <span class="html-italic">SI</span><sub>a2-2</sub> under Cas 7D or Mode 8H.</p>
Full article ">Figure 7
<p>Corresponding relationship between the pole pair number and the electric cycle number of DFISs.</p>
Full article ">Figure 8
<p><span class="html-italic">SI</span><sub>m</sub> under different loading conditions of the machine.</p>
Full article ">Figure 9
<p><span class="html-italic">SI</span><sub>m</sub> under various speeds of the machine.</p>
Full article ">Figure 10
<p>The process for automatically diagnosing demagnetization faults using three PNNs.</p>
Full article ">Figure 11
<p>Architecture of the PNN.</p>
Full article ">Figure 12
<p>The training results of the three PNNs. (<b>a</b>) The first PNN. (<b>b</b>) The second PNN. (<b>c</b>) The third PNN.</p>
Full article ">Figure 13
<p>The testing results of the three PNNs. (<b>a</b>) The first PNN. (<b>b</b>) The second PNN. (<b>c</b>) The third PNN.</p>
Full article ">Figure 14
<p>Experimental setup.</p>
Full article ">Figure 15
<p>The experimental results of the voltage of TYC1, TYC2, and SI<sub>m</sub> under healthy conditions. (<b>a</b>) The voltage waveform of TYC1 and TYC2. (<b>b</b>) The voltage waveform of <span class="html-italic">SI</span><sub>m</sub>.</p>
Full article ">Figure 16
<p>The experimental results of residual voltage of under fault3. (<b>a</b>) The residual voltage of TYC1, TYC2, and TYC3. (<b>b</b>) The residual voltage waveform of <span class="html-italic">S</span><sub>m</sub> and <span class="html-italic">S</span><sub>a1</sub>.</p>
Full article ">
15 pages, 7442 KiB  
Article
Simulation of Groundwater-Level Behavior in Southeast Region of Korea Induced by 2016 Gyeong-Ju Earthquake Using 2D Hydro-Mechanical Coupled Bonded Particle Modeling
by Hyunjin Cho, Se-Yeong Hamm, Jeoung Seok Yoon, Soo-Gin Kim and Jae-Yeol Cheong
Appl. Sci. 2024, 14(24), 11939; https://doi.org/10.3390/app142411939 - 20 Dec 2024
Viewed by 499
Abstract
This study examines the hydrogeological response to the 12 September 2016 Gyeong-Ju earthquake (ML 5.8) in the southeastern region of the Korean Peninsula. Using 2D hydro-mechanical coupled bonded particle modeling, we simulated the dynamic fault rupture process to analyze stress redistribution and its [...] Read more.
This study examines the hydrogeological response to the 12 September 2016 Gyeong-Ju earthquake (ML 5.8) in the southeastern region of the Korean Peninsula. Using 2D hydro-mechanical coupled bonded particle modeling, we simulated the dynamic fault rupture process to analyze stress redistribution and its impact on pore pressure and groundwater levels (GWLs). The results indicated that compressional areas correlated strongly with pore pressure increases and GWL rises, while extensional areas showed decreases in both. Observations from the groundwater monitoring Well 5 at Gyeong-Ju San-Nae and Well 8 at Gyeong-Ju Cheon-Buk, located approximately 15 km from the earthquake’s epicenter, aligned well with the model’s predictions and interpretation, providing validation for the simulation. These findings highlight the capability of hydro-mechanical models to capture fault-induced hydrological responses and offer valuable insights into the interplay between seismic activity and groundwater systems. Full article
(This article belongs to the Special Issue Progress and Challenges of Rock Engineering)
Show Figures

Figure 1

Figure 1
<p>Major faults distributed in the southeastern region of the Korean Peninsula. The modeled area is shown by the red rectangle. The black star is the location of the 2016 GJEQ. MYF: Mir-Yang Fault, USF: Ul-San Fault, MRF: Mo-Ryang Fault, YSF: Yang-San Fault, DRF: Dong-Rae Fault.</p>
Full article ">Figure 2
<p>Locations of GWL monitoring wells distributed within the studied area centered around the hypocenter of the GJEQ. The co-seismic GWL changes monitored at the wells (numbers and names are given) induced by the GJEQ are represented by triangles (level rise) and the amount of level rise is depicted by the color. The black star is the location of the GJEQ. Temporal evolution of the GWL at the numbered wells are provided in the <a href="#app1-applsci-14-11939" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 3
<p>Workflow for fault dynamic rupture simulation using PFC2D v7.00.161.</p>
Full article ">Figure 4
<p>Faults represented by smooth joint contact model and the orientations of the in situ horizontal stresses shown by the stress tensor. Black star is the hypocenter of the 2016 GJEQ.</p>
Full article ">Figure 5
<p>(<b>a</b>) Pore–pipe network implemented in a bonded particle assembly and (<b>b</b>) description of how pore pressure exerts force (multiplication of pore pressure <span class="html-italic">P<sub>f</sub></span>, length <span class="html-italic">l<sub>i</sub></span>, and model’s out-of-plane direction thickness <span class="html-italic">t</span>) to the surrounding particles (hydro-mechanical coupled effect).</p>
Full article ">Figure 6
<p>Algorithm of hydro-mechanical coupling processes in PFC2D modeling (a solid line arrow means iterative process within a single code, while a broken line arrow means iterative communications between different codes).</p>
Full article ">Figure 7
<p>(<b>a</b>) Particle displacement vectors at 5 s after start of fault dynamic rupture. The displacement vectors show right-lateral slip of the ruptured fault. (<b>b</b>) Spatial distribution of the changes in the maximum horizontal stress due to fault rupture. The black star denotes the location of the 2016 GJEQ.</p>
Full article ">Figure 8
<p>Spatial and temporal evolution of the particle velocity (m/s) field at (<b>a</b>) 1.2 s, (<b>b</b>) 4.8 s, and (<b>c</b>) 10.8 s after start of fault rupture. Figures at different time instances are provided in the <a href="#app1-applsci-14-11939" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 9
<p>Spatial distribution of pore pressure at simulation time of (<b>a</b>) 1.2 s, (<b>b</b>) 4.8 s, and (<b>c</b>) 10.8 s after start of fault rupture. Figures at different time instances are provided in the <a href="#app1-applsci-14-11939" class="html-app">Supplementary Materials</a>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Pore pressure evolution simulated at the location of Well 5 in the model, corresponding to the location of Station 5 (Gyeong-Ju San-Nae) and (<b>b</b>) GWL changes monitored at Station 5 Gyeong-Ju San-Nae. The horizontal axis is the elapsed days since 1 January 2016. The red arrow indicates the onset of the GWL changes induced by the GJEQ.</p>
Full article ">Figure 11
<p>(<b>a</b>) Pore pressure evolution simulated at the location of Well 8 in the model, corresponding to the location of Station 8 (Gyeong-Ju Cheon-Buk) and (<b>b</b>) GWL changes monitored at Station 8 Gyeong-Ju Cheon-Buk. The horizontal axis is the elapsed days since 1 January 2016. The red arrow indicates the onset of level changes induced by the GJEQ.</p>
Full article ">
22 pages, 4466 KiB  
Article
Assessment of the Geological Storage Potential and Suitability of CO2 in the Deep Saline Aquifers in the Northwest Plain of Shandong Province, China
by Shihao Wang, Hailong Tian, Xincun Zhao, Yan Yan, Xunchang Yang and Xuepeng Wang
Energies 2024, 17(24), 6387; https://doi.org/10.3390/en17246387 - 19 Dec 2024
Viewed by 391
Abstract
Carbon capture and storage (CCS) technology is a crucial and effective tool to achieve China’s dual carbon goals. The primary locations suitable for underground CO2 storage include depleted oil and gas reservoirs, deep saline aquifers, and deep unmineable coal seams. Among these, [...] Read more.
Carbon capture and storage (CCS) technology is a crucial and effective tool to achieve China’s dual carbon goals. The primary locations suitable for underground CO2 storage include depleted oil and gas reservoirs, deep saline aquifers, and deep unmineable coal seams. Among these, deep saline aquifers are widely distributed in most of the world’s sedimentary basins, and they offer significant advantages—such as substantial storage capacity, well-established technology, high safety standards, and cost effectiveness—making them crucial geological reservoirs for carbon dioxide storage. In comparison to foreign countries’ projects on CO2 capture, utilization, and storage (CCUS) technology, China’s initiatives have been implemented more recently, and no research has been conducted on the geological storage of CO2 in the deep saline aquifers within the study area. In this study, we systematically analyzed the key factors for the geological storage of CO2 in saline reservoirs within the northwest plain of Shandong Province: the Paleogene Shahejie Formation saline aquifer, and the lower reservoir of the Minghuazhen Formation saline aquifer located east of the Zhanhua–Lijin–Dongying line. The CO2 geological storage potential of these aquifers was assessed using the evaluation methodology of the United States Department of Energy, yielding a result of 30.355 billion tons. An evaluation index system of CO2 geological storage suitability was established. Evaluation indices for regions in the study area were assigned according to this evaluation index, and the score and grade of each unit were obtained. The results indicated that the Huimin latent fault depression, Dongying latent fault depression, Dezhou latent fault depression, and Dongming–Shenxian latent fault depression are suitable prospective areas for CO2 geological storage in the saline aquifers of Shandong Province’s northwest plain. Full article
(This article belongs to the Section D: Energy Storage and Application)
Show Figures

Figure 1

Figure 1
<p>Map of the study area.</p>
Full article ">Figure 2
<p>Geological structure map of northwest Shandong.</p>
Full article ">Figure 3
<p>CO<sub>2</sub> geological sequestration potential technology and economic resource pyramid model.</p>
Full article ">Figure 4
<p>Pyramid diagram illustrating the assessment stages of China’s potential for CO<sub>2</sub> geological storage.</p>
Full article ">Figure 5
<p>Hierarchical model diagram.</p>
Full article ">Figure 6
<p>Index hierarchy model for evaluating CO<sub>2</sub> geological storage suitability.</p>
Full article ">Figure 7
<p>Histogram of weights of each indicator.</p>
Full article ">Figure 8
<p>Suitability zoning map of CO<sub>2</sub> geological storage for each assessment unit.</p>
Full article ">
18 pages, 21691 KiB  
Article
Knowledge Graph-Based In-Context Learning for Advanced Fault Diagnosis in Sensor Networks
by Xin Xie, Junbo Wang, Yu Han and Wenjuan Li
Sensors 2024, 24(24), 8086; https://doi.org/10.3390/s24248086 - 18 Dec 2024
Viewed by 441
Abstract
This paper introduces a novel approach for enhancing fault diagnosis in industrial equipment systems through the application of sensor network-driven knowledge graph-based in-context learning (KG-ICL). By focusing on the critical role of sensor data in detecting and isolating faults, we construct a domain-specific [...] Read more.
This paper introduces a novel approach for enhancing fault diagnosis in industrial equipment systems through the application of sensor network-driven knowledge graph-based in-context learning (KG-ICL). By focusing on the critical role of sensor data in detecting and isolating faults, we construct a domain-specific knowledge graph (DSKG) that encapsulates expert knowledge relevant to industrial equipment. Utilizing a long-length entity similarity (LES) measure, we retrieve relevant information from the DSKG. Our method leverages large language models (LLMs) to conduct causal analysis on textual data related to equipment faults derived from sensor networks, thereby significantly enhancing the accuracy and efficiency of fault diagnosis. This paper details a series of experiments that validate the effectiveness of the KG-ICL method in accurately diagnosing fault causes and locations of industrial equipment systems. By leveraging LLMs and structured knowledge, our approach offers a robust tool for condition monitoring and fault management, thereby improving the reliability and efficiency of operations in industrial sectors. Full article
(This article belongs to the Special Issue Fault Diagnosis in Sensor Network-Based Systems)
Show Figures

Figure 1

Figure 1
<p>Process of KG-ICL. This process encompasses KG construction (Step one), relevant subgraph matching (Step two), and prompt generation (Step three). Ultimately, we utilize the generated prompts to assist LLMs in arriving at accurate conclusions for fault cause analysis (Step four).</p>
Full article ">Figure 2
<p>Schema layer of the DSKG. The schema layer is divided into two main parts: an equipment-related sublayer and a fault-related sublayer.</p>
Full article ">Figure 3
<p>Part of a DSKG. Different colors in the figure represent different types of nodes, a total of 1800 nodes (<span class="html-italic">★</span>), the number of different nodes is also indicated in the legend.</p>
Full article ">Figure 4
<p>Subgraph retrieval. Subgraph retrieval is carried out mainly by means of a subgraph retriever. The process includes (<b>a</b>) data input, (<b>b</b>) KG query, (<b>c</b>) data filtering, and (<b>d</b>) subgraph generation. The white nodes represent entities within the knowledge graph, while the yellow and green nodes denote task-relevant nodes. Among these, the yellow nodes exhibit a higher degree of relevance to the current task.</p>
Full article ">Figure 5
<p>Subgraph retriever. (<b>a</b>) Coarse-grained classification is performed on the long-length entities extracted from the input text to determine whether they contain fault information. (<b>b</b>) Next, we perform word segmentation to obtain equipment-related information and (<b>c</b>) identify potentially relevant parts of the DSKG. (<b>d</b>) Finally, the relevant subgraph is obtained based on similarity to retrieve the information of interest.</p>
Full article ">Figure 6
<p>Prompt template.</p>
Full article ">Figure 7
<p>Accuracy in terms of Position Correct and Cause Correct outcomes (k = 10).</p>
Full article ">Figure 8
<p>The confusion matrix of the results (k = 10). A more intense blue color signifies a higher proportion.</p>
Full article ">Figure 8 Cont.
<p>The confusion matrix of the results (k = 10). A more intense blue color signifies a higher proportion.</p>
Full article ">
19 pages, 5844 KiB  
Article
The Detection of an Inter-Turn Short-Circuit Fault in a Brushless Permanent Magnet Motor with Different Winding Configurations
by Mariusz Korkosz, Jan Prokop and Karol Ryłło
Energies 2024, 17(24), 6379; https://doi.org/10.3390/en17246379 - 18 Dec 2024
Viewed by 386
Abstract
This article is about the detection of a partial inter-turn short-circuit fault in a brushless motor with permanent magnets (BLPMM). The detection of a single inter-turn short circuit is a difficult issue. The authors of this article tested the sensitivity of the tested [...] Read more.
This article is about the detection of a partial inter-turn short-circuit fault in a brushless motor with permanent magnets (BLPMM). The detection of a single inter-turn short circuit is a difficult issue. The authors of this article tested the sensitivity of the tested powertrain damage detection method. The diagnostic method developed for BLPMM allows for any configuration of the motor winding. A number of analysed configurations have been applied for the combined star–delta connection (YΔ). For the combined star–delta connection Y/Δ, the problem of partial short circuit at two locations was analysed. In the first case, this was the short-circuit winding phenomena in the star part (SC1). In the second case, the short circuit was connected in part to the delta (SC2). A mathematical model has been developed that takes into account both the type of connection and the chance of a partial short circuit of the coil. Based on numerical calculations, the sensitivity of diagnostic methods is designated for both cases. Furthermore, the impact of partial short circuits on motor performance was also examined. The effect of a partial inter-turn short-circuit fault on motor parameters was also determined. Laboratory verification was carried out. Full article
(This article belongs to the Special Issue Reliability and Condition Monitoring of Electric Motors and Drives)
Show Figures

Figure 1

Figure 1
<p>View (<b>a</b>) hybrid drive and (<b>b</b>) cross-section of BLPMM.</p>
Full article ">Figure 2
<p>Scheme (<b>a</b>) power converter, (<b>b</b>) winding connections.</p>
Full article ">Figure 3
<p>Flowchart of the proposed monitoring and diagnostic method.</p>
Full article ">Figure 4
<p>Laboratory test rig.</p>
Full article ">Figure 5
<p>RMS SC current vs. speed.</p>
Full article ">Figure 6
<p>SC current waveform: (<b>a</b>) stage of SC1, (<b>b</b>) stage of SC2.</p>
Full article ">Figure 7
<p>RMS SC current vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 8
<p>RMS current vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>: (<b>a</b>) stage of SC1, (<b>b</b>) stage of SC2.</p>
Full article ">Figure 9
<p>Electromagnetic torque <span class="html-italic">T</span><sub>eav</sub> vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 10
<p>RMS of diagnostic signal <span class="html-italic">u</span><sub>0</sub> vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 11
<p>Magnitude of diagnostic signal vs. number of shorted turns: (<b>a</b>) stage of SC1, (<b>b</b>) stage of SC2.</p>
Full article ">Figure 12
<p>Percentage of the magnitude <span class="html-italic">f</span><sub>1</sub>/<span class="html-italic">f</span><sub>3</sub> vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 13
<p>Harmonic amplitudes <span class="html-italic">f</span><sub>1</sub> and <span class="html-italic">f</span><sub>3</sub> vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 14
<p>The difference in amplitude <span class="html-italic">f</span><sub>1</sub> and <span class="html-italic">f</span><sub>3</sub> vs. number of shorted turns <span class="html-italic">N</span><sub>sc</sub>.</p>
Full article ">Figure 15
<p>Current waveforms—SYM laboratory test.</p>
Full article ">Figure 16
<p>Current waveforms—SC1 laboratory test.</p>
Full article ">Figure 17
<p>Current waveforms—SC2 laboratory test.</p>
Full article ">Figure 18
<p>FFT of diagnostic signal: (<b>a</b>) linear scale, (<b>b</b>) decibel scale.</p>
Full article ">
Back to TopTop