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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (214)

Search Parameters:
Keywords = rough terrain

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
32 pages, 11087 KiB  
Article
Path Planning and Motion Control of Robot Dog Through Rough Terrain Based on Vision Navigation
by Tianxiang Chen, Yipeng Huangfu, Sutthiphong Srigrarom and Boo Cheong Khoo
Sensors 2024, 24(22), 7306; https://doi.org/10.3390/s24227306 - 15 Nov 2024
Viewed by 639
Abstract
This article delineates the enhancement of an autonomous navigation and obstacle avoidance system for a quadruped robot dog. Part one of this paper presents the integration of a sophisticated multi-level dynamic control framework, utilizing Model Predictive Control (MPC) and Whole-Body Control (WBC) from [...] Read more.
This article delineates the enhancement of an autonomous navigation and obstacle avoidance system for a quadruped robot dog. Part one of this paper presents the integration of a sophisticated multi-level dynamic control framework, utilizing Model Predictive Control (MPC) and Whole-Body Control (WBC) from MIT Cheetah. The system employs an Intel RealSense D435i depth camera for depth vision-based navigation, which enables high-fidelity 3D environmental mapping and real-time path planning. A significant innovation is the customization of the EGO-Planner to optimize trajectory planning in dynamically changing terrains, coupled with the implementation of a multi-body dynamics model that significantly improves the robot’s stability and maneuverability across various surfaces. The experimental results show that the RGB-D system exhibits superior velocity stability and trajectory accuracy to the SLAM system, with a 20% reduction in the cumulative velocity error and a 10% improvement in path tracking precision. The experimental results also show that the RGB-D system achieves smoother navigation, requiring 15% fewer iterations for path planning, and a 30% faster success rate recovery in challenging environments. The successful application of these technologies in simulated urban disaster scenarios suggests promising future applications in emergency response and complex urban environments. Part two of this paper presents the development of a robust path planning algorithm for a robot dog on a rough terrain based on attached binocular vision navigation. We use a commercial-of-the-shelf (COTS) robot dog. An optical CCD binocular vision dynamic tracking system is used to provide environment information. Likewise, the pose and posture of the robot dog are obtained from the robot’s own sensors, and a kinematics model is established. Then, a binocular vision tracking method is developed to determine the optimal path, provide a proposal (commands to actuators) of the position and posture of the bionic robot, and achieve stable motion on tough terrains. The terrain is assumed to be a gentle uneven terrain to begin with and subsequently proceeds to a more rough surface. This work consists of four steps: (1) pose and position data are acquired from the robot dog’s own inertial sensors, (2) terrain and environment information is input from onboard cameras, (3) information is fused (integrated), and (4) path planning and motion control proposals are made. Ultimately, this work provides a robust framework for future developments in the vision-based navigation and control of quadruped robots, offering potential solutions for navigating complex and dynamic terrains. Full article
Show Figures

Figure 1

Figure 1
<p>Simplified box model of the Lite3P quadruped robotic dog.</p>
Full article ">Figure 2
<p>Internal sensor arrangement of the quadruped robotic dog.</p>
Full article ">Figure 3
<p>Dynamic control flowchart.</p>
Full article ">Figure 4
<p>MPC flowchart.</p>
Full article ">Figure 5
<p>WBC flowchart [<a href="#B30-sensors-24-07306" class="html-bibr">30</a>].</p>
Full article ">Figure 6
<p>Robot coordinates and joint point settings [<a href="#B30-sensors-24-07306" class="html-bibr">30</a>].</p>
Full article ">Figure 7
<p>Intel D435i and velodyne LIDAR.</p>
Full article ">Figure 8
<p>ICP diagram.</p>
Full article ">Figure 9
<p>Comparison of before and after modifying the perception region.</p>
Full article ">Figure 10
<p>Point cloud processing flowchart.</p>
Full article ">Figure 11
<p>{p, v} generation: (<b>a</b>) the creation of {p, v} pairs for collision points; (<b>b</b>) the process of generating anchor points and repulsive vectors for dynamic obstacle avoidance [<a href="#B41-sensors-24-07306" class="html-bibr">41</a>].</p>
Full article ">Figure 12
<p>Overall framework of 2D EGO-Planner.</p>
Full article ">Figure 13
<p>Robot initialization and control process in Gazebo simulation: (<b>a</b>) Gazebo environment creation, (<b>b</b>) robot model import, (<b>c</b>) torque balance mode activation, and (<b>d</b>) robot stepping and rotation in simulation.</p>
Full article ">Figure 14
<p>Joint rotational angles of FL and RL legs.</p>
Full article ">Figure 15
<p>Joint angular velocities of FL and RL legs.</p>
Full article ">Figure 16
<p>Torque applied to FL and RL joints during the gait cycle.</p>
Full article ">Figure 17
<p>The robot navigating in a simple environment using a camera.</p>
Full article ">Figure 18
<p>The robot navigating in a complex environment using a camera.</p>
Full article ">Figure 19
<p>A 2D trajectory showing start and goal positions, obstacles, and rough path.</p>
Full article ">Figure 20
<p>Initial environment setup.</p>
Full article ">Figure 21
<p>The robot starts navigating in a simple environment with a static obstacle (brown box).</p>
Full article ">Figure 22
<p>Dynamic Obstacle 1 introduced: the robot detects a new obstacle and recalculates its path.</p>
Full article ">Figure 23
<p>Dynamic Obstacle 2 introduced: after avoiding the first obstacle, a second obstacle is introduced and detected by the planner.</p>
Full article ">Figure 24
<p>Approaching the target: the robot adjusts its path to approach the target point as the distance shortens.</p>
Full article ">Figure 25
<p>Reaching the target: the robot completes its path and reaches the designated target point.</p>
Full article ">Figure 26
<p>Real-time B-spline trajectory updates in response to dynamic obstacles. Set 1 (orange) shows the initial path avoiding static obstacles. When the first dynamic obstacle is detected, the EGO-Planner updates the path (Set 2, blue) using local optimization. A second obstacle prompts another adjustment (Set 3, green), guiding the robot smoothly towards the target as trajectory updates become more frequent.</p>
Full article ">Figure 27
<p>The robot navigating a simple environment using SLAM.</p>
Full article ">Figure 28
<p>The robot navigating a complex environment using SLAM.</p>
Full article ">Figure 29
<p>A 2D trajectory showing start and goal positions, obstacles, and the planned path in a complex environment using SLAM.</p>
Full article ">Figure 30
<p>Navigation based on RGB-D camera.</p>
Full article ">Figure 31
<p>Navigation based on SLAM.</p>
Full article ">Figure 32
<p>Velocity deviation based on RGB-D camera.</p>
Full article ">Figure 33
<p>Velocity deviation based on SLAM.</p>
Full article ">Figure 34
<p>Cumulative average iterations.</p>
Full article ">Figure 35
<p>Cumulative success rate.</p>
Full article ">
13 pages, 3510 KiB  
Article
Laboratory Validation of 3D Model and Investigating Its Application to Wind Turbine Noise Propagation over Rough Ground
by James Naylor and Qin Qin
Wind 2024, 4(4), 363-375; https://doi.org/10.3390/wind4040018 - 7 Nov 2024
Viewed by 376
Abstract
In an investigation into how wind turbine noise interacts with the surrounding terrain, its propagation over rough ground is simulated using a parabolic equation code using a modified effective impedance model, which characterizes the effects of a three-dimensional, rigid roughness within a relatively [...] Read more.
In an investigation into how wind turbine noise interacts with the surrounding terrain, its propagation over rough ground is simulated using a parabolic equation code using a modified effective impedance model, which characterizes the effects of a three-dimensional, rigid roughness within a relatively long wavelength limit (ka1). The model is validated by comparison to experiments conducted within an anechoic chamber wherein different source–receiver geometries are considered. The relative sound pressure level spectra from the parabolic equation code using the modified effective impedance model highlight a sensitivity to the roughness parameters. At a low frequency and far distance, the relative sound pressure level decreased as the roughness coverage increased. A difference of 4.9 dB has been reported. The simulations highlight how the roughness shifts the ground effect dips, resulting in the sound level at the distance of 2 km being altered. However, only the monochromatic wave has been discussed. Further work on broadband noise is desirable. Furthermore, due to the long wavelength limit, only a portion of audible wind turbine noise can be investigated. Full article
Show Figures

Figure 1

Figure 1
<p>Roughness types. (<b>a</b>) Closely packed; (<b>b</b>) Medium packed; (<b>c</b>) Sparsely packed.</p>
Full article ">Figure 2
<p>Imaginary part of the normalized effective impedance calculated using the Attenborough–Tolstoy impedance model for various roughness types.</p>
Full article ">Figure 3
<p>Diagram of experimental setup.</p>
Full article ">Figure 4
<p>Comparison of the measured excess attenuation to the analytical and parabolic equation theory using the Attenborough–Tolstoy effective impedance model. (<b>a</b>) Closely packed roughness; (<b>b</b>) Medium packed roughness; (<b>c</b>) Sparsely packed roughness. For all figures (<b>a</b>–<b>c</b>), the sound source height, receiver height and horizontal distance between the two were 100 mm, 100 mm and 400 mm, respectively, for situation 1.</p>
Full article ">Figure 4 Cont.
<p>Comparison of the measured excess attenuation to the analytical and parabolic equation theory using the Attenborough–Tolstoy effective impedance model. (<b>a</b>) Closely packed roughness; (<b>b</b>) Medium packed roughness; (<b>c</b>) Sparsely packed roughness. For all figures (<b>a</b>–<b>c</b>), the sound source height, receiver height and horizontal distance between the two were 100 mm, 100 mm and 400 mm, respectively, for situation 1.</p>
Full article ">Figure 5
<p>Comparison of the measured relative sound pressure level, Rel. SPL, to the analytical and parabolic equation theory using the Attenborough–Tolstoy effective impedance model for a closely packed roughness. The sound source height, receiver height and horizontal distance between the two were 200 mm, 20 mm, and 100 mm, respectively, for situation 2.</p>
Full article ">Figure 6
<p>Simplified diagram of the scenario in mind for the parabolic equation method simulations.</p>
Full article ">Figure 7
<p>Relative sound pressure level, Rel. SPL, against propagation range, measured at 2 m height. Calculated for various roughness types using the parabolic equation method.</p>
Full article ">Figure 8
<p>Relative sound pressure level, Rel. SPL, map close to ground level. Calculated using the parabolic equation method. (<b>a</b>) Sparsely packed roughness; (<b>b</b>) Medium packed roughness; (<b>c</b>) Closely packed roughness.</p>
Full article ">Figure 8 Cont.
<p>Relative sound pressure level, Rel. SPL, map close to ground level. Calculated using the parabolic equation method. (<b>a</b>) Sparsely packed roughness; (<b>b</b>) Medium packed roughness; (<b>c</b>) Closely packed roughness.</p>
Full article ">
23 pages, 21957 KiB  
Article
Terrain Analysis According to Multiscale Surface Roughness in the Taklimakan Desert
by Sebastiano Trevisani and Peter L. Guth
Land 2024, 13(11), 1843; https://doi.org/10.3390/land13111843 - 5 Nov 2024
Viewed by 510
Abstract
Surface roughness, interpreted in the wide sense of surface texture, is a generic term referring to a variety of aspects and scales of spatial variability of surfaces. The analysis of solid earth surface roughness is useful for understanding, characterizing, and monitoring geomorphic factors [...] Read more.
Surface roughness, interpreted in the wide sense of surface texture, is a generic term referring to a variety of aspects and scales of spatial variability of surfaces. The analysis of solid earth surface roughness is useful for understanding, characterizing, and monitoring geomorphic factors at multiple spatiotemporal scales. The different geomorphic features characterizing a landscape exhibit specific characteristics and scales of surface texture. The capability to selectively analyze specific roughness metrics at multiple spatial scales represents a key tool in geomorphometric analysis. This research presents a simplified geostatistical approach for the multiscale analysis of surface roughness, or of image texture in the case of images, that is highly informative and interpretable. The implemented approach is able to describe two main aspects of short-range surface roughness: omnidirectional roughness and roughness anisotropy. Adopting simple upscaling approaches, it is possible to perform a multiscale analysis of roughness. An overview of the information extraction potential of the approach is shown for the analysis of a portion of the Taklimakan desert (China) using a 30 m resolution DEM derived from the Copernicus Glo-30 DSM. The multiscale roughness indexes are used as input features for unsupervised and supervised learning tasks. The approach can be refined both from the perspective of the multiscale analysis as well as in relation to the surface roughness indexes considered. However, even in its present, simplified form, it can find direct applications in relation to multiple contexts and research topics. Full article
(This article belongs to the Section Land, Soil and Water)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Reprojected COP DEM (30 m resolution, UTM F44) of the area of interest overlaid on the hillshade.</p>
Full article ">Figure 2
<p>Sentinel-2 true color RGB image (bands 4, 3, and 2) of the study area, with the main dune morphologies labeled.</p>
Full article ">Figure 3
<p>Main dune morphologies in the study area, visualized using Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>). From top to bottom: network/transverse dunes, longitudinal and transverse dunes, and dome-shaped dunes.</p>
Full article ">Figure 4
<p>Mixed morphologies in the area of interest, visualized using Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>). From top to bottom: outcropping bedrock with shadow and linear dunes, fluvial morphology, and a flat area with minor dune morphologies.</p>
Full article ">Figure 5
<p>RA direction, where the RA strength is higher than 0.3, overlaid on the hillshade (<b>a</b>) and the residual DEM (<b>b</b>) calculated for level L2.</p>
Full article ">Figure 6
<p>Omnidirectional short-range roughness (m) for the different resolutions. Different color scales for each diagram.</p>
Full article ">Figure 7
<p>Roughness anisotropy strength at different resolutions. Different color scales for each diagram.</p>
Full article ">Figure 8
<p>RGB image (each band normalized) of 3 omnidirectional roughness indexes computed at different resolutions (B = L1; G = L2; R = L4). Despite the high correlation of the three indexes, they differentiate very well the morphological features of the area. For example, they markedly highlight the characteristic smoothness of interdune areas of the longitudinal dunes south of the mountain ridge.</p>
Full article ">Figure 9
<p>RGB image (each band normalized) of 3 anisotropy strength roughness indexes computed at different resolutions (B = L1; G = L4; R = L16). In the dune fields north of the mountains, long-wavelength anisotropic features prevail; in contrast, for the southern longitudinal dunes, shorter anisotropic features (L4) are highlighted.</p>
Full article ">Figure 10
<p>Landscape clustered according to multiscale surface roughness indexes. The cluster centers in terms of OR and RA are described in <a href="#land-13-01843-f011" class="html-fig">Figure 11</a>.</p>
Full article ">Figure 11
<p>Cluster centers of the 7 classes resulting from K-means clustering for OR and RA at the different levels.</p>
Full article ">Figure 12
<p>MRI clustering results in the area of the northern dune field, characterized by network and transverse dunes. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 13
<p>MRI clustering results in the area of the southern longitudinal dune fields. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 14
<p>MRI clustering results in the area with fluvial morphology, outcropping bedrock, and dome dune fields. Clustering results (<b>d</b>), Sentinel-2 imagery (<b>a</b>), hillshade (<b>b</b>), and residual DEM (<b>c</b>).</p>
Full article ">Figure 15
<p>Manual classification of crest lines (<b>a</b>) for large dunes using visual analysis of slope (<b>b</b>), profile curvature (<b>c</b>), and residual DEM (<b>d</b>). Crest lines are associated with high positive profile curvature, strongly positive residual DEM, and low slope. These locations are then located in areas in which the neighborhood is characterized by an abrupt variation in the selected geomorphometric derivatives.</p>
Full article ">Figure 16
<p>Probability of observing a crest obtained by means of RF considering the GDs integrated with the MRIs (<b>a</b>) and only the five GDs (<b>b</b>) to obtain details of the study area, which is located on the western mountain ridge. The RF model integrating the MRIs provides a more focused prediction of crest lines of large dunes. In (<b>c</b>), the prediction of the crest lines of the two RF models is compared. Pixels with a probability higher than 0.8 have been classified as crests. The transparent color is where both models predicted a not-crest pixel, green is where both models predicted a crest, and red and blue are where, respectively, only RF GDs and RF GDs + MRIs predicted a crest.</p>
Full article ">Figure 17
<p>Variables’ importance in the two RF models according to the mean decrease in the Gini index ((<b>a</b>), RF based on GDs; (<b>b</b>), RF based on GDs integrated with MRIs).</p>
Full article ">Figure 18
<p>Prediction of crest lines with the RF model based on GDs and MRIs of an unseen area ((<b>c</b>), green box) external to the one with reference data used for training and testing ((<b>c</b>), red box). The reference crest lines (<b>a</b>) have been manually digitized by means of visual analysis of the profile curvature, the residual DEM, and the slope; the predicted crest lines have been derived as crests of all of the pixels with a probability above 0.8. The predicted crest lines are compared with the reference data (<b>b</b>). Green pixels are correctly classified as crests; red and blue pixels are incorrectly classified, respectively, as crests and not crests.</p>
Full article ">
25 pages, 6213 KiB  
Article
Simulation of the Neutral Atmospheric Flow Using Multiscale Modeling: Comparative Studies for SimpleFoam and Fluent Solver
by Zihan Zhao, Lingxiao Tang and Yiqing Xiao
Atmosphere 2024, 15(10), 1259; https://doi.org/10.3390/atmos15101259 - 21 Oct 2024
Viewed by 463
Abstract
The reproduced planetary boundary layer (PBL) wind is commonly applied in downscaled simulations using commercial CFD codes with Reynolds-averaged Navier–Stokes (RANS) turbulence modeling. When using the turbulent inlets calculated by numerical weather prediction models (NWP), adjustments of the turbulence eddy viscosity closures and [...] Read more.
The reproduced planetary boundary layer (PBL) wind is commonly applied in downscaled simulations using commercial CFD codes with Reynolds-averaged Navier–Stokes (RANS) turbulence modeling. When using the turbulent inlets calculated by numerical weather prediction models (NWP), adjustments of the turbulence eddy viscosity closures and wall function formulations are concerned with maintaining the fully developed wind profiles specified at the inlet of CFD domains. The impact of these related configurations is worth discussing in engineering applications, especially when commercial codes restrict the internal modifications. This study evaluates the numerical performances of open-source OpenFOAM 2.3.0 and commercial Fluent 17.2 codes as supplementary scientific comparisons. This contribution focuses on the modified turbulence closures to incorporate turbulent profiles produced from mesoscale PBL parameterizations and the modified wall treatments relating to the roughness length. The near-ground flow features are evaluated by selecting the flat terrains and the classical Askervein benchmark case. The improvement in near-ground wind flow under the downscaled framework shows satisfactory performance in the open-source CFD platform. This contributes to engineers realizing the micro-siting of wind turbines and quantifying terrain-induced speed-up phenomena under the scope of wind-resistant design. Full article
(This article belongs to the Section Atmospheric Techniques, Instruments, and Modeling)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the full-scale CFD domain with flat terrain. Five positions of wind profiles (D0~D2500) are extracted along the streamwise direction, where the number represents the straight-line distance from the inlet. The purple gridlines represent the mesh grid. (<b>b</b>) Details of the surface-grid resolution in the streamwise direction.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) are the structured grids on Askervein Hill case with central refinements close to measurements (Line A-A, AA-AA, RS and HT). Note that the positive X and Y axes represent the east and north directions, respectively. The green color gridlines represent the surface mesh grid. (<b>c</b>) Details of the XZ plane grid along with the hilltop HT. The red gridlines represent the spatial mesh grid, and blue background color is the working space during grid generation.</p>
Full article ">Figure 3
<p>The concerned parameters in the <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>−</mo> <mi>ε</mi> </mrow> </semantics></math> turbulence solver using the multiscale modeling.</p>
Full article ">Figure 4
<p>Summary of the inlet conditions used in the flat terrain case. (<b>a</b>) Wind magnitude profile, (<b>b</b>) turbulent kinetic energy profile, (<b>c</b>) dissipation profile.</p>
Full article ">Figure 5
<p>Summary of the three-hour averaged wind profiles extracted from WRF solutions on the Askervein TU-03B field campaign. Twenty-eight positions are specified on the west (W-WRF) and south (S-WRF) of CFD lateral boundaries. (<b>a</b>–<b>c</b>) Wind components and directions. (<b>d</b>,<b>e</b>) Turbulent kinetic energy and dissipation profiles from WRF PBL parameterizations.</p>
Full article ">Figure 6
<p>Illustration of offline data transfer between mesoscale coarse profiles data (<b>left</b>) and CFD fine-grained boundaries (<b>right</b>). The different colors represent the interpolated variables.</p>
Full article ">Figure 7
<p>Comparison of inlet and outlet wind velocity profiles (<b>a</b>,<b>d</b>), turbulent kinetic profiles (<b>b</b>,<b>e</b>), and turbulent kinetic energy dissipation (<b>c</b>,<b>f</b>). The model constants and wall functions are configured with the standard treatments listed in <a href="#atmosphere-15-01259-t003" class="html-table">Table 3</a>. The different color lines represent the position extracted in <a href="#atmosphere-15-01259-f001" class="html-fig">Figure 1</a>. (FL-Fluent code, OF-OpenFOAM code).</p>
Full article ">Figure 7 Cont.
<p>Comparison of inlet and outlet wind velocity profiles (<b>a</b>,<b>d</b>), turbulent kinetic profiles (<b>b</b>,<b>e</b>), and turbulent kinetic energy dissipation (<b>c</b>,<b>f</b>). The model constants and wall functions are configured with the standard treatments listed in <a href="#atmosphere-15-01259-t003" class="html-table">Table 3</a>. The different color lines represent the position extracted in <a href="#atmosphere-15-01259-f001" class="html-fig">Figure 1</a>. (FL-Fluent code, OF-OpenFOAM code).</p>
Full article ">Figure 8
<p>Comparison of inlet and outlet wind velocity profiles (<b>a</b>,<b>d</b>), turbulent kinetic profiles (<b>b</b>,<b>e</b>), and turbulent kinetic energy dissipation (<b>c</b>,<b>f</b>). The numerical cases are configured with the standard model constants and modified wall functions as listed in <a href="#atmosphere-15-01259-t003" class="html-table">Table 3</a>. The different color lines represent the position extracted in <a href="#atmosphere-15-01259-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 9
<p>Comparison of inlet and outlet wind velocity profiles (<b>a</b>,<b>b</b>), turbulent kinetic profiles (<b>c</b>,<b>d</b>), and turbulent kinetic energy dissipation (<b>e</b>,<b>f</b>). The numerical cases are configured with the modified model constants and the standard wall functions, as listed in <a href="#atmosphere-15-01259-t003" class="html-table">Table 3</a>. The different color lines represent the position extracted in <a href="#atmosphere-15-01259-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 9 Cont.
<p>Comparison of inlet and outlet wind velocity profiles (<b>a</b>,<b>b</b>), turbulent kinetic profiles (<b>c</b>,<b>d</b>), and turbulent kinetic energy dissipation (<b>e</b>,<b>f</b>). The numerical cases are configured with the modified model constants and the standard wall functions, as listed in <a href="#atmosphere-15-01259-t003" class="html-table">Table 3</a>. The different color lines represent the position extracted in <a href="#atmosphere-15-01259-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 10
<p>Profiles of terrain-induced wind speed-up <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> </semantics></math> and turbulent kinetic energy <math display="inline"><semantics> <mi>k</mi> </semantics></math>. Numerical cases are configured with the standard model constants (SC) and the wall function (SW). (<b>a</b>,<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> </semantics></math> along line A-A and line AA-AA. (<b>c</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math> along lines A-A. The horizontally distributed profiles are 10 m AGL in height. (<b>d</b>) The vertical <math display="inline"><semantics> <mi>k</mi> </semantics></math> profile in the hilltop HT site. D-HT and D-CP represent the distance from HT and CP positions.</p>
Full article ">Figure 11
<p>Contours of the horizontal wind velocities (<b>a</b>,<b>b</b>) and turbulent kinetic energy (<b>c</b>,<b>d</b>) distributed with 10 m AGL height. The black and white circles in (<b>a</b>,<b>b</b>) denote the maximum and minimum wind speeds in the hilltop and leeward regions, respectively. The numerical cases are configured with the standard model constants (SC) and near-wall treatments (SW).</p>
Full article ">Figure 12
<p>The modified wall functions (MW) are configured in the numerical cases. (<b>a</b>,<b>b</b>) are <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> </semantics></math> along line A-A and line AA-AA. (<b>c</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math> along lines A-A. The horizontally distributed profiles are 10 m AGL in height. (<b>d</b>) The vertical <math display="inline"><semantics> <mi>k</mi> </semantics></math> profile in the hilltop HT site.</p>
Full article ">Figure 13
<p>The horizontally distributed contours of wind velocities (<b>a</b>,<b>b</b>) and turbulent kinetic energy (<b>c</b>,<b>d</b>) are configured using the modified near-wall functions.</p>
Full article ">Figure 14
<p>The modified turbulence model constants (MC) are configured in the numerical case. (<b>a</b>,<b>b</b>) are <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>S</mi> </mrow> </semantics></math> along line A-A and line AA-AA. (<b>c</b>) <math display="inline"><semantics> <mi>k</mi> </semantics></math> along lines A-A. The horizontally distributed profiles are 10 m AGL in height. (<b>d</b>) The vertical <math display="inline"><semantics> <mi>k</mi> </semantics></math> profile in the hilltop HT site.</p>
Full article ">Figure 15
<p>As for <a href="#atmosphere-15-01259-f011" class="html-fig">Figure 11</a>, the horizontally distributed contours of wind velocities (<b>a</b>,<b>b</b>) and turbulent kinetic energy (<b>c</b>,<b>d</b>) are configured using the modified turbulence closures.</p>
Full article ">Figure 15 Cont.
<p>As for <a href="#atmosphere-15-01259-f011" class="html-fig">Figure 11</a>, the horizontally distributed contours of wind velocities (<b>a</b>,<b>b</b>) and turbulent kinetic energy (<b>c</b>,<b>d</b>) are configured using the modified turbulence closures.</p>
Full article ">
31 pages, 33586 KiB  
Article
A Fuzzy Pure Pursuit for Autonomous UGVs Based on Model Predictive Control and Whole-Body Motion Control
by Yaoyu Sui, Zhong Yang, Haoze Zhuo, Yulong You, Wenqiang Que and Naifeng He
Drones 2024, 8(10), 554; https://doi.org/10.3390/drones8100554 - 6 Oct 2024
Viewed by 723
Abstract
In this paper, we propose an adaptive fuzzy pure pursuit trajectory tracking algorithm for autonomous unmanned ground vehicles (UGVs), addressing the challenges of accurate and stable navigation in complex environments. Traditional pure pursuit methods with fixed look-ahead distances struggle to maintain precision in [...] Read more.
In this paper, we propose an adaptive fuzzy pure pursuit trajectory tracking algorithm for autonomous unmanned ground vehicles (UGVs), addressing the challenges of accurate and stable navigation in complex environments. Traditional pure pursuit methods with fixed look-ahead distances struggle to maintain precision in dynamic and uneven terrains. Our approach uniquely integrates a fuzzy control algorithm that allows for real-time adjustments of the look-ahead distance based on environmental feedback, thereby enhancing tracking accuracy and smoothness. Additionally, we combine this with model predictive control (MPC) and whole-body motion control (WBC), where MPC forecasts future states and optimally adjusts control actions, while WBC ensures coordinated motion of the UGV, maintaining balance and stability, especially in rough terrains. This integration not only improves responsiveness to changing conditions but also enables dynamic balance adjustments during movement. The proposed algorithm was validated through simulations in Gazebo and real-world experiments on physical platforms. In real-world tests, our algorithm reduced the average trajectory tracking error by 45% and the standard deviation by nearly 50%, significantly improving stability and accuracy compared to traditional methods. Full article
(This article belongs to the Special Issue Advances in Guidance, Navigation, and Control)
Show Figures

Figure 1

Figure 1
<p>The quadruped robot used in the physical experiments is about to take on the challenge of tracking an S-shaped curve in complex outdoor terrain.</p>
Full article ">Figure 2
<p>Simplified model of quadruped robot.</p>
Full article ">Figure 3
<p>Pure pursuit for quadruped robots.</p>
Full article ">Figure 4
<p>Fuzzy logic diagram.</p>
Full article ">Figure 5
<p>Fuzzy inputs and outputs.</p>
Full article ">Figure 6
<p>The control quantity surfaces for <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>k</mi> <mi>v</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>k</mi> <mi>w</mi> </msub> </mrow> </semantics></math> are listed as (<b>a</b>) the control quantity surfaces for <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>k</mi> <mi>v</mi> </msub> </mrow> </semantics></math> and (<b>b</b>) the control quantity surfaces for <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>k</mi> <mi>w</mi> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>MPC–WBC framework diagram.</p>
Full article ">Figure 8
<p>Simulation in Gazebo.</p>
Full article ">Figure 9
<p>Quadruped robot used for simulation.</p>
Full article ">Figure 10
<p>Ground undulation 3 cm test (walk in line).</p>
Full article ">Figure 11
<p>Robot desired trajectory vs. actual trajectory (3 cm, line).</p>
Full article ">Figure 12
<p>Position errors, angular error, and velocities during robot walking (3 cm, line).</p>
Full article ">Figure 13
<p>Three joint torques of one leg (3 cm, line).</p>
Full article ">Figure 14
<p>(<b>a</b>) Value of <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>v</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>w</mi> </msub> </mrow> </semantics></math>; (<b>b</b>) value of <math display="inline"><semantics> <mi>l</mi> </semantics></math>.</p>
Full article ">Figure 15
<p>Ground undulation 6 cm test (work in line).</p>
Full article ">Figure 16
<p>Robot desired trajectory vs. actual trajectory (6 cm, line).</p>
Full article ">Figure 17
<p>Position errors, angular error, and velocities during robot walking (6 cm, line).</p>
Full article ">Figure 18
<p>Three joint torques of one leg (6 cm, line).</p>
Full article ">Figure 19
<p>(<b>a</b>) Value of <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>v</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>w</mi> </msub> </mrow> </semantics></math>; (<b>b</b>) value of <math display="inline"><semantics> <mi>l</mi> </semantics></math>.</p>
Full article ">Figure 20
<p>Ground undulation 6 cm test (walk in circle).</p>
Full article ">Figure 21
<p>Robot desired trajectory vs. actual trajectory (6 cm, circle).</p>
Full article ">Figure 22
<p>Position errors, angular error, and velocities during robot walking (6 cm, circle).</p>
Full article ">Figure 23
<p>Three joint torques of one leg (6 cm, circle).</p>
Full article ">Figure 24
<p>(<b>a</b>) Value of <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>v</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>w</mi> </msub> </mrow> </semantics></math>; (<b>b</b>) value of <math display="inline"><semantics> <mi>l</mi> </semantics></math>.</p>
Full article ">Figure 25
<p>USLGO1.</p>
Full article ">Figure 26
<p>Demonstration of experimental robot walking routes in complex outdoor environments. (<b>a</b>–<b>d</b>) represent the 4 parts of the S-curve respectively.</p>
Full article ">Figure 27
<p>Comparison of real-world robot trajectories with different algorithms.</p>
Full article ">Figure 28
<p>Position errors, angular error, and velocities during robot walking (in real situation).</p>
Full article ">Figure 29
<p>(<b>a</b>) Three joint torques of one leg (in real situation); (<b>b</b>) value of <math display="inline"><semantics> <mi>l</mi> </semantics></math>.</p>
Full article ">
16 pages, 34354 KiB  
Article
Autonomous Vehicles Traversability Mapping Fusing Semantic–Geometric in Off-Road Navigation
by Bo Zhang, Weili Chen, Chaoming Xu, Jinshi Qiu and Shiyu Chen
Drones 2024, 8(9), 496; https://doi.org/10.3390/drones8090496 - 18 Sep 2024
Viewed by 1078
Abstract
This paper proposes an evaluating and mapping methodology of terrain traversability for off-road navigation of autonomous vehicles in unstructured environments. Terrain features are extracted from RGB images and 3D point clouds to create a traversal cost map. The cost map is then employed [...] Read more.
This paper proposes an evaluating and mapping methodology of terrain traversability for off-road navigation of autonomous vehicles in unstructured environments. Terrain features are extracted from RGB images and 3D point clouds to create a traversal cost map. The cost map is then employed to plan safe trajectories. Bayesian generalized kernel inference is employed to assess unknown grid attributes due to the sparse raw point cloud data. A Kalman filter also creates density local elevation maps in real time by fusing multiframe information. Consequently, the terrain semantic mapping procedure considers the uncertainty of semantic segmentation and the impact of sensor noise. A Bayesian filter is used to update the surface semantic information in a probabilistic manner. Ultimately, the elevation map is utilized to extract geometric characteristics, which are then integrated with the probabilistic semantic map. This combined map is then used in conjunction with the extended motion primitive planner to plan the most effective trajectory. The experimental results demonstrate that the autonomous vehicles obtain a success rate enhancement ranging from 4.4% to 13.6% and a decrease in trajectory roughness ranging from 5.1% to 35.8% when compared with the most developed outdoor navigation algorithms. Additionally, the autonomous vehicles maintain a terrain surface selection accuracy of over 85% during the navigation process. Full article
Show Figures

Figure 1

Figure 1
<p>Navigation of an autonomous vehicle in unstructured wild environments with different terrains and obstacles.</p>
Full article ">Figure 2
<p>Autonomous navigation framework overview. The algorithm evaluates terrain traversability using RGB images and 3D point clouds. The traversability map guarantees safe and smooth path planning.</p>
Full article ">Figure 3
<p>Semantic cost projection. Some examples of projection: green represents easily traversable areas such as asphalt and concrete, yellow represents rough surfaces such as grass and sand, orange represents rocks, and blue represents obstacles.</p>
Full article ">Figure 4
<p>Comparison of obstacle identification results of the three methods in two real-world scenarios. In the scenario, Case A (first row) has two hidden obstacle areas (two puddle areas 1, 2); Case B (second row) has three false obstacles (three individual tall grassy areas 3, 4, and 5). (<b>a</b>) RGB image inputs (corresponding to the two images in the first column) for both cases; (<b>b</b>) semantic segmentation for both cases; (<b>c</b>) geometric gridmaps for both cases; (<b>d</b>) fused passivity maps for both cases.</p>
Full article ">Figure 5
<p>Results of evaluating the terrain for geometry only (top) and geometric-semantic fusion (bottom). (<b>a</b>) Scenario 1, (<b>b</b>) Scenario 2, and (<b>c</b>) Scenario 3. The areas that the geometric method gets wrong are shown in blue, and the areas where the geometry is neglected are displayed in red.</p>
Full article ">Figure 6
<p>Autonomous vehicle trajectories when navigating in four various unstructured scenarios using proposed method (black), RSPMP (blue), PUTN (red), BGK (orange), and DWA (yellow). (<b>a</b>) Scenario 1; (<b>b</b>) Scenario 2; (<b>c</b>) Scenario 3; (<b>d</b>) Scenario 4. It can be observed that proposed method allows the autonomous vehicle to navigate on smooth, low-cost surfaces and maintains a short trajectory.</p>
Full article ">Figure 7
<p>Quantitative comparison of navigation performance between various methods. (<b>a</b>) Scenario 1; (<b>b</b>) Scenario 2; (<b>c</b>) Scenario 3; (<b>d</b>) Scenario 4.</p>
Full article ">
20 pages, 9948 KiB  
Article
Traversability Analysis and Path Planning for Autonomous Wheeled Vehicles on Rigid Terrains
by Nan Wang, Xiang Li, Zhe Suo, Jiuchen Fan, Jixin Wang and Dongxuan Xie
Drones 2024, 8(9), 419; https://doi.org/10.3390/drones8090419 - 23 Aug 2024
Viewed by 764
Abstract
Autonomous vehicles play a crucial role in three-dimensional transportation systems and have been extensively investigated and implemented in mining and other fields. However, the diverse and intricate terrain characteristics present challenges to vehicle traversability, including complex geometric features such as slope, harsh physical [...] Read more.
Autonomous vehicles play a crucial role in three-dimensional transportation systems and have been extensively investigated and implemented in mining and other fields. However, the diverse and intricate terrain characteristics present challenges to vehicle traversability, including complex geometric features such as slope, harsh physical parameters such as friction and roughness, and irregular obstacles. The current research on traversability analysis primarily emphasizes the processing of perceptual information, with limited consideration for vehicle performance and state parameters, thereby restricting their applicability in path planning. A framework of traversability analysis and path planning methods for autonomous wheeled vehicles on rigid terrains is proposed in this paper for better traversability costs and less redundancy in path planning. The traversability boundary conditions are established first based on terrain and vehicle characteristics using theoretical methods to determine the traversable areas. Then, the traversability cost map for the traversable areas is obtained through simulation and segmented linear regression analysis. Afterward, the TV-Hybrid A* algorithm is proposed by redefining the path cost functions of the Hybrid A* algorithm through the simulation data and neural network method to generate a more cost-effective path. Finally, the path generated by the TV-Hybrid A* algorithm is validated and compared with that of the A* and Hybrid A* algorithms in simulations, demonstrating a slightly better traversability cost for the former. Full article
Show Figures

Figure 1

Figure 1
<p>An example of rigid terrain composed of compacted soil.</p>
Full article ">Figure 2
<p>The proposed framework of traversability analysis and path planning methods for autonomous wheeled vehicles on rigid terrains.</p>
Full article ">Figure 3
<p>The relationship among vehicle speed, vehicle load, and terrain parameters: (<b>a</b>) The maximum average driving speed of vehicles under different loads on varying terrain slopes with favorable friction and roughness conditions; (<b>b</b>) The maximum average driving speed of vehicles under different loads on varying friction conditions of flat terrain with favorable roughness conditions; (<b>c</b>) The maximum average driving speed of vehicles under different loads on varying roughness conditions of flat terrain with favorable friction conditions. The <span class="html-italic">x</span>-axis values 1–8 correspond to the roughness categories A–H.</p>
Full article ">Figure 3 Cont.
<p>The relationship among vehicle speed, vehicle load, and terrain parameters: (<b>a</b>) The maximum average driving speed of vehicles under different loads on varying terrain slopes with favorable friction and roughness conditions; (<b>b</b>) The maximum average driving speed of vehicles under different loads on varying friction conditions of flat terrain with favorable roughness conditions; (<b>c</b>) The maximum average driving speed of vehicles under different loads on varying roughness conditions of flat terrain with favorable friction conditions. The <span class="html-italic">x</span>-axis values 1–8 correspond to the roughness categories A–H.</p>
Full article ">Figure 4
<p>A three-layer neural network to analyze the impact of terrain parameters on path costs.</p>
Full article ">Figure 5
<p>The simulation validation framework for traversability analysis and path planning methods on rigid terrains.</p>
Full article ">Figure 6
<p>The initial environmental information for the simulation validation: (<b>a</b>) The terrain DEM model; (<b>b</b>) The terrain height map obtained from the DEM model; (<b>c</b>) The terrain gradient map obtained from the DEM model; (<b>d</b>) The terrain physical distribution information of friction marked in gray with roughness marked with filling; (<b>e</b>) The terrain obstacles with height and type information.</p>
Full article ">Figure 6 Cont.
<p>The initial environmental information for the simulation validation: (<b>a</b>) The terrain DEM model; (<b>b</b>) The terrain height map obtained from the DEM model; (<b>c</b>) The terrain gradient map obtained from the DEM model; (<b>d</b>) The terrain physical distribution information of friction marked in gray with roughness marked with filling; (<b>e</b>) The terrain obstacles with height and type information.</p>
Full article ">Figure 6 Cont.
<p>The initial environmental information for the simulation validation: (<b>a</b>) The terrain DEM model; (<b>b</b>) The terrain height map obtained from the DEM model; (<b>c</b>) The terrain gradient map obtained from the DEM model; (<b>d</b>) The terrain physical distribution information of friction marked in gray with roughness marked with filling; (<b>e</b>) The terrain obstacles with height and type information.</p>
Full article ">Figure 7
<p>The simulation model of the articulated chassis vehicle on the designed terrain.</p>
Full article ">Figure 8
<p>Traversability boundary map with black areas indicating impassable areas.</p>
Full article ">Figure 9
<p>The traversability cost maps with the impassable areas marked in black: (<b>a</b>) Height cost map; (<b>b</b>) Slope cost map along <span class="html-italic">x</span>-direction gradient; (<b>c</b>) Slope cost map along <span class="html-italic">y</span>-direction gradient; (<b>d</b>) Friction cost map; (<b>e</b>) Roughness cost map; (<b>f</b>) Obstacles cost map.</p>
Full article ">Figure 9 Cont.
<p>The traversability cost maps with the impassable areas marked in black: (<b>a</b>) Height cost map; (<b>b</b>) Slope cost map along <span class="html-italic">x</span>-direction gradient; (<b>c</b>) Slope cost map along <span class="html-italic">y</span>-direction gradient; (<b>d</b>) Friction cost map; (<b>e</b>) Roughness cost map; (<b>f</b>) Obstacles cost map.</p>
Full article ">Figure 10
<p>The paths generated by the Hybrid A* and TV-Hybrid A* planners.</p>
Full article ">
10 pages, 1022 KiB  
Technical Note
A Simple Path to the Small Perturbation Method for Scattering from Slightly Rough Dielectric Surfaces
by Antonio Iodice and Pasquale Imperatore
Remote Sens. 2024, 16(16), 3035; https://doi.org/10.3390/rs16163035 - 18 Aug 2024
Viewed by 823
Abstract
We propose a perturbative method to compute electromagnetic scattering from slightly rough dielectric surfaces, which leads to the same result as the usual Small Perturbation Method (SPM) in a surprisingly simple way. The proposed method is based on three pillars: the volumetric perturbative [...] Read more.
We propose a perturbative method to compute electromagnetic scattering from slightly rough dielectric surfaces, which leads to the same result as the usual Small Perturbation Method (SPM) in a surprisingly simple way. The proposed method is based on three pillars: the volumetric perturbative approach, the reciprocity theorem, and a proper approximation of the electric field within the perturbation volume, that we name Internal Field Approximation (IFA). The proposed new mathematical derivation of the SPM turns out to be much simpler and more concise than the classical one. In addition, being based on a volumetric perturbation approach, it has the potential of dealing in future with surface and volume scattering within a unitary framework, which is useful in modelling scattering from, e.g., vegetated soil, snow-covered terrain, and inhomogeneous soils. Therefore, although the presented result is mainly theoretical, it can have important applications in remote sensing. Full article
(This article belongs to the Section Engineering Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Geometry of the problem.</p>
Full article ">Figure 2
<p>Decomposition scheme for the perturbation volume.</p>
Full article ">
24 pages, 24217 KiB  
Article
Evaluating the Impact of DEM Spatial Resolution on 3D Rockfall Simulation in GIS Environment
by Maria P. Kakavas, Paolo Frattini, Alberto Previati and Konstantinos G. Nikolakopoulos
Geosciences 2024, 14(8), 200; https://doi.org/10.3390/geosciences14080200 - 29 Jul 2024
Cited by 1 | Viewed by 853
Abstract
Rockfalls are natural geological phenomena characterized by the abrupt detachment and freefall descent of rock fragments from steep slopes. These events exhibit considerable variability in scale, velocity, and trajectory, influenced by the geological composition of the slope, the topography, and other environmental conditions. [...] Read more.
Rockfalls are natural geological phenomena characterized by the abrupt detachment and freefall descent of rock fragments from steep slopes. These events exhibit considerable variability in scale, velocity, and trajectory, influenced by the geological composition of the slope, the topography, and other environmental conditions. By employing advanced modeling techniques and terrain analysis, researchers aim to predict and control rockfall hazards to prevent casualties and protect properties in areas at risk. In this study, two rockfall events in the villages of Myloi and Platiana of Ilia prefecture were examined. The research was conducted by means of HY-STONE software, which performs 3D numerical modeling of the motion of non-interacting blocks. To perform this modeling, input files require the processing of base maps and datasets in a GIS environment. Stochastic modeling and 3D descriptions of slope topography, based on Digital Elevation Models (DEMs) without spatial resolution limitations, ensure multiscale analysis capabilities. Considering this capability, seven freely available DEMs, derived from various sources, were applied in HY-STONE with the scope of performing a large number of multiparametric analyses and selecting the most appropriate and efficient DEM for the software requirements. All the necessary data for the multiparametric analyses were generated within a GIS environment, utilizing either the same restitution coefficients and rolling friction coefficient or varying ones. The results indicate that finer-resolution DEMs capture detailed terrain features, enabling the precise identification of rockfall source areas and an accurate depiction of the kinetic energy distribution. Further, the results show that a correct application of the model to different DEMs requires a specific parametrization to account for the different roughness of the models. Full article
(This article belongs to the Special Issue Earth Observation by GNSS and GIS Techniques)
Show Figures

Figure 1

Figure 1
<p>Pictures of the rockfall events in the Myloi and Platiana region in relation to the Hellenic region.</p>
Full article ">Figure 2
<p>Pictures of the rockfall in the Myloi region illustrate the final position of rock blocks. Red arrows indicate the final positions of the rock masses at the conclusion of the rockfall event.</p>
Full article ">Figure 3
<p>Picture taken from the Platiana region after the rockfall event. The slope picture is taken from the Greek Cadastral (2008).</p>
Full article ">Figure 4
<p>Reforestation procedures in the slope next to the Platiana village. The slope pictures are taken from the Greek Cadastral.</p>
Full article ">Figure 5
<p>Myloi area before (<b>left</b> image) and after (<b>right</b> image) removing the vegetation through Cloud Compare Software.</p>
Full article ">Figure 6
<p>HY-STONE results for the site of Myloi using parameters (E<sub>N</sub>, E<sub>T</sub>, and A<sub>T</sub>) calibrated on the UAV DEM. The transit frequencies of blocks are shown for (<b>a</b>) UAV DEM, (<b>b</b>) Greek Cadastral DEM, (<b>c</b>) ALOS AW3D30 DEM, (<b>d</b>) ASTER GDEM, and (<b>e</b>) SRTM30 DEM.</p>
Full article ">Figure 7
<p>HY-STONE results for the site of Myloi using parameters (E<sub>N</sub>, E<sub>T</sub>, and A<sub>T</sub>) calibrated on the UAV DEM. The maximum translation kinetic energies are shown for (<b>a</b>) UAV DEM, (<b>b</b>) Greek Cadastral DEM, (<b>c</b>) ALOS AW3D30 DEM, (<b>d</b>) ASTER GDEM, and (<b>e</b>) SRTM30 DEM.</p>
Full article ">Figure 8
<p>Slope profiles from (<b>a</b>) UAV DSE, (<b>b</b>) Greek Cadastral DEM, (<b>c</b>) ALOS AW3D30 DEM, (<b>d</b>) ASTER GDEM, and (<b>e</b>) SRTM30 DEM.</p>
Full article ">Figure 9
<p>HY-STONE results, in terms of transit frequency for the site of Platiana. (<b>a</b>) Greek Cadastral DEM, (<b>b</b>) ALOS AW3D30 DEM, (<b>c</b>) ASTER GDEM, (<b>d</b>) SRTM30 DEM, (<b>e</b>) SRTM90 DEM, and (<b>f</b>) TanDEM_X.</p>
Full article ">Figure 10
<p>HY-STONE results in terms of the maximum translation kinetic energy for the site of Platiana. (<b>a</b>) Greek Cadastral DEM, (<b>b</b>) ALOS AW3D30 DEM, (<b>c</b>) ASTER GDEM, (<b>d</b>) SRTM30 DEM, (<b>e</b>) SRTM90 DEM, and (<b>f</b>) TanDEM_X.</p>
Full article ">Figure 11
<p>Slope profiles from (<b>a</b>) Greek Cadastral DEM, (<b>b</b>) ALOS AW3D30 DEM, (<b>c</b>) ASTER GDEM, (<b>d</b>) SRTM30 DEM, (<b>e</b>) SRTM90 DEM, and (<b>f</b>) TanDEM_X.</p>
Full article ">Figure 12
<p>HY-STONE results by using the Greek Cadastral DEM with the optimal coefficients shown in <a href="#geosciences-14-00200-t003" class="html-table">Table 3</a> for the site of Myloi: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 13
<p>HY-STONE results by using the ALOS AW3D30 DEM with optimal coefficients shown in <a href="#geosciences-14-00200-t004" class="html-table">Table 4</a> for the site of Platiana: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 14
<p>HY-STONE results by using the ASTER GDEM with the optimal coefficients shown in <a href="#geosciences-14-00200-t005" class="html-table">Table 5</a> for the site of Platiana: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 15
<p>HY-STONE results by using the SRTM30 DEM with the optimal coefficients shown in <a href="#geosciences-14-00200-t006" class="html-table">Table 6</a> for the site of Platiana: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 16
<p>HY-STONE results by using the SRTM90 DEM with the optimal coefficients shown in <a href="#geosciences-14-00200-t007" class="html-table">Table 7</a> for the site of Platiana: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 17
<p>HY-STONE results by using the TanDEM_X with the optimal coefficients shown in <a href="#geosciences-14-00200-t008" class="html-table">Table 8</a> for the site of Platiana: (<b>a</b>) the transit frequency and (<b>b</b>) the maximum translation kinetic energy.</p>
Full article ">Figure 18
<p>The effect of spatial resolution in the normal (E<sub>N</sub>) and tangential (E<sub>T</sub>) restitutions and the rolling friction (A<sub>T</sub>) coefficients. For 30 m and 90 m, three and two DEMS are available, respectively.</p>
Full article ">
17 pages, 5892 KiB  
Article
Improved A* Algorithm for Mobile Robots under Rough Terrain Based on Ground Trafficability Model and Ground Ruggedness Model
by Zhiguang Liu, Song Guo, Fei Yu, Jianhong Hao and Peng Zhang
Sensors 2024, 24(15), 4884; https://doi.org/10.3390/s24154884 - 27 Jul 2024
Cited by 1 | Viewed by 897
Abstract
Considering that the existing path planning algorithms for mobile robots under rugged terrain do not consider the ground flatness and the lack of optimality, which leads to the instability of the center of mass of the mobile robot, this paper proposes an improved [...] Read more.
Considering that the existing path planning algorithms for mobile robots under rugged terrain do not consider the ground flatness and the lack of optimality, which leads to the instability of the center of mass of the mobile robot, this paper proposes an improved A* algorithm for mobile robots under rugged terrain based on the ground accessibility model and the ground ruggedness model. Firstly, the ground accessibility and ruggedness models are established based on the elevation map, expressing the ground flatness. Secondly, the elevation cost function that can obtain the optimal path is designed based on the two types of models combined with the characteristics of the A* algorithm, and the continuous cost function is established by connecting with the original distance cost function, which avoids the center-of-mass instability caused by the non-optimal path. Finally, the effectiveness of the improved algorithm is verified by simulation and experiment. The results show that compared with the existing commonly used path planning algorithms under rugged terrain, the enhanced algorithm improves the smoothness of paths and the optimization degree of paths in the path planning process under rough terrain. Full article
(This article belongs to the Topic Advances in Mobile Robotics Navigation, 2nd Volume)
Show Figures

Figure 1

Figure 1
<p>The autonomous navigation system framework. The mobile robot establishes an elevation map and a target node based on the goal pose and point cloud information, after which a new path is generated by the path planner module and tracked by the controller until the mobile robot reaches the node where the goal pose is located.</p>
Full article ">Figure 2
<p>Ground trafficability model. The elevation difference model represents the difference between the child and parent nodes, where (<span class="html-italic">x</span><sub>1</sub>, <span class="html-italic">y</span><sub>1</sub>, <span class="html-italic">z</span><sub>1</sub>) and (<span class="html-italic">x</span><sub>2</sub>, <span class="html-italic">y</span><sub>2</sub>, <span class="html-italic">z</span><sub>2</sub>) represent the parent and child node coordinates, respectively. The slope model represents the slope between the child node and the parent node, and (<span class="html-italic">x</span><sub>1</sub>, <span class="html-italic">y</span><sub>1</sub>, <span class="html-italic">z</span><sub>1</sub>) and (<span class="html-italic">x</span><sub>3</sub>, <span class="html-italic">y</span><sub>3</sub>, <span class="html-italic">z</span><sub>3</sub>) represent the parent and child node coordinates, respectively.</p>
Full article ">Figure 3
<p>Ground ruggedness model. The ground ruggedness model describes the degree of fluctuation of the search ground. The child node is (<span class="html-italic">x<sub>i</sub></span>, <span class="html-italic">y<sub>j</sub></span>, <span class="html-italic">z<sub>i</sub></span><sub>,<span class="html-italic">j</span></sub>).</p>
Full article ">Figure 4
<p>Ackermann steering mobile robot platforms. The robot is built with 3D LiDAR for sensing and is processed and controlled by Nvidia Jetson TX2 running Ubuntu 18.04 and ROS 1.12.17.</p>
Full article ">Figure 5
<p>MATLAB rough terrain simulation map.</p>
Full article ">Figure 6
<p>Paths formed by the mobile robot path planning in MATLAB rough terrain simulation map: (<b>a</b>) shows the path formed by the OA* algorithm; (<b>b</b>) shows the path formed by the AOEA* algorithm; and (<b>c</b>) shows the path formed by the IA* algorithm.</p>
Full article ">Figure 7
<p>Paths are formed by the mobile robot path planning in the Gazebo rough terrain simulation, in which the red path is the path created by the AOEA* algorithm, and the yellow path is the path made by the IA* algorithm. Arrows and dots represent the start and end points of paths.</p>
Full article ">Figure 8
<p>The <span class="html-italic">x</span>-, <span class="html-italic">y</span>-, and <span class="html-italic">z</span>-axis angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation: (<b>a</b>) shows the <span class="html-italic">x</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation; (<b>b</b>) shows the <span class="html-italic">y</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation; and (<b>c</b>) shows the <span class="html-italic">z</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation.</p>
Full article ">Figure 9
<p>Paths are formed by the mobile robot path planning in the real-life environment, in which the yellow path is the path created by the AOEA* algorithm, and the blue path is the path made by the IA* algorithm. Arrows and dots represent the start and end points of paths.</p>
Full article ">Figure 10
<p>The <span class="html-italic">x</span>-, <span class="html-italic">y</span>-, and <span class="html-italic">z</span>-axis angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation: (<b>a</b>) shows the <span class="html-italic">x</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation; (<b>b</b>) shows the <span class="html-italic">y</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation; and (<b>c</b>) shows the <span class="html-italic">z</span> angles of each node of the mobile robot to the ground coordinate system for the paths planned by the two algorithms in the Gazebo rough terrain simulation.</p>
Full article ">
16 pages, 22837 KiB  
Article
Learning to Walk with Adaptive Feet
by Antonello Scaldaferri, Franco Angelini and Manolo Garabini
Robotics 2024, 13(8), 113; https://doi.org/10.3390/robotics13080113 - 24 Jul 2024
Viewed by 1085
Abstract
In recent years, tasks regarding autonomous mobility favoredthe use of legged robots rather than wheeled ones thanks to their higher mobility on rough and uneven terrains. This comes at the cost of more complex motion planners and controllers to ensure robot stability and [...] Read more.
In recent years, tasks regarding autonomous mobility favoredthe use of legged robots rather than wheeled ones thanks to their higher mobility on rough and uneven terrains. This comes at the cost of more complex motion planners and controllers to ensure robot stability and balance. However, in the case of quadrupedal robots, balancing is simpler than it is for bipeds thanks to their larger support polygons. Until a few years ago, most scientists and engineers addressed the quadrupedal locomotion problem with model-based approaches, which require a great deal of modeling expertise. A new trend is the use of data-driven methods, which seem to be quite promising and have shown great results. These methods do not require any modeling effort, but they suffer from computational limitations dictated by the hardware resources used. However, only the design phase of these algorithms requires large computing resources (controller training); their execution in the operational phase (deployment), takes place in real time on common processors. Moreover, adaptive feet capable of sensing terrain profile information have been designed and have shown great performance. Still, no dynamic locomotion control method has been specifically designed to leverage the advantages and supplementary information provided by this type of adaptive feet. In this work, we investigate the use and evaluate the performance of different end-to-end control policies trained via reinforcement learning algorithms specifically designed and trained to work on quadrupedal robots equipped with passive adaptive feet for their dynamic locomotion control over a diverse set of terrains. We examine how the addition of the haptic perception of the terrain affects the locomotion performance. Full article
(This article belongs to the Special Issue Applications of Neural Networks in Robot Control)
Show Figures

Figure 1

Figure 1
<p>Quadrupedal robot dynamic locomotion with adaptive feet.</p>
Full article ">Figure 2
<p>Differences between hybrid [<a href="#B15-robotics-13-00113" class="html-bibr">15</a>] and end-to-end [<a href="#B14-robotics-13-00113" class="html-bibr">14</a>] approaches.</p>
Full article ">Figure 3
<p>Proposed approximations of the SoftFoot-Q [<a href="#B10-robotics-13-00113" class="html-bibr">10</a>] (<b>a</b>) model: (<b>b</b>) Adaptive Flat Foot (AFF) and (<b>c</b>) Adaptive Open Foot (AOF). The passive DoF rotation axes are highlighted in red.</p>
Full article ">Figure 4
<p>Reinforcement learning environments.</p>
Full article ">Figure 5
<p>Terrain setup for the rough terrain environment.</p>
Full article ">Figure 6
<p>Robot-centric elevation map with sampled point coordinates.</p>
Full article ">Figure 7
<p>Reward trends of the different training sessions. AFF stands for Adaptive Flat Feet (<a href="#robotics-13-00113-f003" class="html-fig">Figure 3</a>b), while AOF stands for Adaptive Open Feet (<a href="#robotics-13-00113-f003" class="html-fig">Figure 3</a>c).</p>
Full article ">Figure 8
<p>Performance of the different robot setups tested for the flat terrain locomotion task.</p>
Full article ">Figure 9
<p>Performance of the different robot setups tested for the rough terrain locomotion task.</p>
Full article ">Figure 10
<p>Performance evaluation when random pushes are applied to the robot base. The disturbance is applied at 8.75 s. The mean and the standard deviation were computed by simulating 1024 robots in parallel, with each one following a random commanded base twist reference.</p>
Full article ">Figure 11
<p>Photo sequence of the P-AFF policy on stairs.</p>
Full article ">Figure 12
<p>Feet contact schedule of the P-AFF policy on stairs.</p>
Full article ">Figure 13
<p>Photo sequence of the PH-AOF policy on a rough slope.</p>
Full article ">Figure 14
<p>Feet contact schedule of the PH-AOF policy on a rough slope.</p>
Full article ">Figure 15
<p>Foot orientation problem.</p>
Full article ">
13 pages, 7462 KiB  
Article
Assessment of Landslide Susceptibility in the Moxi Tableland of China by Using a Combination of Deep-Learning and Factor-Refinement Methods
by Zonghan He, Wenjun Zhang, Jialun Cai, Jing Fan, Haoming Xu, Hui Feng, Xinlong Luo and Zhouhang Wu
Appl. Sci. 2024, 14(12), 5042; https://doi.org/10.3390/app14125042 - 10 Jun 2024
Viewed by 1016
Abstract
Precisely assessing the vulnerability of landslides is essential for effective risk assessment. The findings from such assessments will undoubtedly be in high demand, providing a solid scientific foundation for a range of critical initiatives aimed at disaster prevention and control. In the research, [...] Read more.
Precisely assessing the vulnerability of landslides is essential for effective risk assessment. The findings from such assessments will undoubtedly be in high demand, providing a solid scientific foundation for a range of critical initiatives aimed at disaster prevention and control. In the research, authors set the ancient core district of Sichuan Moxi Ancient Town as the research object; they conduct and give the final result of the geological survey. Fault influences are commonly utilized as key markers for delineating strata in the field of stratigraphy, and the slope distance, slope angle, slope aspect, elevation, terrain undulation, plane curvature, profile curvature, mean curvature, relative elevation, land use type, surface roughness, water influence, distance of the catchment, cumulative water volume, and the Normalized Vegetation Index (NDVI) are used along roads to calculate annual rainfall. With the purpose of the establishment of the evaluation system, there are 17 factors selected in total. Through the landslide-susceptibility assessment by the coupled models of DNN-I-SVM and DNN-I-LR nine factors had been selected; it was found that the Area Under the Curve (AUC) value of the Receiver Operating Characteristic Curve (ROC) was high, and the accuracy of the model is relatively high. The coupler, DNN-I-LR, gives 0.875 of an evaluation accuracy of AUC, higher than DNN-I-SVM, which yielded 0.860. It is necessary to note that, in this region, compared to the DNN-I-SVM model, the DNN-I-LR coupling model has better fitting and prediction abilities. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic location and satellite images of the study area.</p>
Full article ">Figure 2
<p>Distribution of landslide hazards in the study area.</p>
Full article ">Figure 3
<p>Disaster-causing factor map of the research area. (<b>a</b>) Fault buffer-zone distance; (<b>b</b>) slope-gradient factor map; (<b>c</b>) elevation distribution; (<b>d</b>) slope aspect distribution; (<b>e</b>) surface roughness; (<b>f</b>) terrain undulation degree; (<b>g</b>) relative elevation; (<b>h</b>) planar curvature; (<b>i</b>) profile curvature; (<b>j</b>) mean curvature; (<b>k</b>) water system-influence distance; (<b>l</b>) flow direction distribution; (<b>m</b>) catchment-accumulation distribution map of the study area; (<b>n</b>) annual rainfall distribution map of the study area; (<b>o</b>) NDVI factor; (<b>p</b>) road-influence distance; (<b>q</b>) land use type.</p>
Full article ">Figure 3 Cont.
<p>Disaster-causing factor map of the research area. (<b>a</b>) Fault buffer-zone distance; (<b>b</b>) slope-gradient factor map; (<b>c</b>) elevation distribution; (<b>d</b>) slope aspect distribution; (<b>e</b>) surface roughness; (<b>f</b>) terrain undulation degree; (<b>g</b>) relative elevation; (<b>h</b>) planar curvature; (<b>i</b>) profile curvature; (<b>j</b>) mean curvature; (<b>k</b>) water system-influence distance; (<b>l</b>) flow direction distribution; (<b>m</b>) catchment-accumulation distribution map of the study area; (<b>n</b>) annual rainfall distribution map of the study area; (<b>o</b>) NDVI factor; (<b>p</b>) road-influence distance; (<b>q</b>) land use type.</p>
Full article ">Figure 4
<p>Results of the model-based zoning map. (<b>a</b>) Landslide-susceptibility-evaluation result map based on DNN model; (<b>b</b>) landslide-susceptibility zoning results of the DNN-I-SVM coupled model; (<b>c</b>) landslide-susceptibility zoning results of the DNN-I-LR coupled model.</p>
Full article ">Figure 4 Cont.
<p>Results of the model-based zoning map. (<b>a</b>) Landslide-susceptibility-evaluation result map based on DNN model; (<b>b</b>) landslide-susceptibility zoning results of the DNN-I-SVM coupled model; (<b>c</b>) landslide-susceptibility zoning results of the DNN-I-LR coupled model.</p>
Full article ">Figure 5
<p>ROC graph.</p>
Full article ">
21 pages, 5333 KiB  
Article
An Advanced Terrain Vegetation Signal Detection Approach for Forest Structural Parameters Estimation Using ICESat-2 Data
by Yifan Li, Xin Shen and Lin Cao
Remote Sens. 2024, 16(11), 1822; https://doi.org/10.3390/rs16111822 - 21 May 2024
Viewed by 1118
Abstract
Accurate forest structural parameters (such as forest height and canopy cover) support forest carbon monitoring, sustainable forest management, and the implementation of silvicultural practices. The Ice, Cloud, and land Elevation Satellite-2 (ICESat-2), which is a spaceborne Light Detection and Ranging (LiDAR) satellite, offers [...] Read more.
Accurate forest structural parameters (such as forest height and canopy cover) support forest carbon monitoring, sustainable forest management, and the implementation of silvicultural practices. The Ice, Cloud, and land Elevation Satellite-2 (ICESat-2), which is a spaceborne Light Detection and Ranging (LiDAR) satellite, offers significant potential for acquiring precise and extensive information on forest structural parameters. However, the ICESat-2 ATL08 product is significantly influenced by the geographical environment and forest characteristics, maintaining considerable potential for enhancing the accuracy of forest height estimation. Meanwhile, it does not focus on providing canopy cover data. To acquire accurate forest structural parameters, the Terrain Signal Neural Network (TSNN) framework was proposed, integrating Computer Vision (CV), Ordering Points to Identify the Clustering Structure (OPTICS), and deep learning. It encompassed an advanced approach for detecting terrain vegetation signals and constructing deep learning models for estimating forest structural parameters using ICESat-2 ATL03 raw data. First, the ATL03 footprints were visualized as Profile Raster Images of Footprints (PRIF), implementing image binarization through adaptive thresholding and median filtering denoising to detect the terrain. Second, the rough denoising buffers were created based on the terrain, combining with the OPTICS clustering and Gaussian denoising algorithms to recognize the terrain vegetation signal footprints. Finally, deep learning models (convolutional neural network (CNN), ResNet50, and EfficientNetB3) were constructed, training standardized PRIF to estimate forest structural parameters (including forest height and canopy cover). The results indicated that the TSNN achieved high accuracy in terrain detection (coefficient of determination (R2) = 0.97) and terrain vegetation signal recognition (F-score = 0.72). The EfficientNetB3 model achieved the highest accuracy in forest height estimation (R2 = 0.88, relative Root Mean Squared Error (rRMSE) = 13.5%), while the CNN model achieved the highest accuracy in canopy cover estimation (R2 = 0.80, rRMSE = 18.5%). Our results have significantly enhanced the accuracy of acquiring ICESat-2 forest structural parameters, while also proposing an original approach combining CV and deep learning for utilizing spaceborne LiDAR data. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Summary of the coverage and distribution of ICESat-2 ATL03 data in the study area. (<b>a</b>) Guangxi is located in southwest China and shares borders with Yunnan, Guizhou, Hunan, and Guangdong provinces; (<b>b</b>) Nanning is located in the center of Guangxi, and the Gaofeng Forest Farm is positioned in the middle of Nanning; (<b>c</b>) The topographical layout and data coverage of ICESat-2 ATL03 data (from October 2018 to October 2020) within the study area.</p>
Full article ">Figure 2
<p>The framework of the advanced terrain vegetation signal detection approach for forest structural parameters estimation using ICESat-2 data (PRIF—Profile Raster Images of Footprints; OPTICS—Ordering Points to Identify the Clustering Structure).</p>
Full article ">Figure 3
<p>The illustration of the improved OPTICS algorithm. OPTICS was utilized to recognize the terrain vegetation signal footprints in this study. In this figure, the yellow scatter points represent noise footprints outside the buffer of DEM trend line, while the black scatters represent rough denoising footprints. The gray dots are the footprints covered by the searching area. OPTICS was employed to establish a search area (red ellipse areas with the red-colored searching center footprints) and cluster the footprints based on their spatial distribution correlation.</p>
Full article ">Figure 4
<p>The standardization of Profile Raster Images of Footprints (PRIF) at different forest structural parameters values. (<b>a</b>) The standardization of PRIF for forest height; (<b>b</b>) The standardization of PRIF for forest height. Denoised PRIFs and standardized PRIFs are in pairs derived from the same ATL03 segment in the 2 rows of each sub-figure. Each column signifies the progressive increase in forest structural parameters values from left to right.</p>
Full article ">Figure 5
<p>The construction of CNN model in this study. The output of each layer is represented by the gray rectangle, while the window size for each convolution step is indicated by the orange rectangle. Additionally, the pink square pyramid illustrates the connection relationship between the convolution window and the corresponding pixels in the output layer.</p>
Full article ">Figure 6
<p>The stem structure of EfficientNet.</p>
Full article ">Figure 7
<p>The illustration of terrain detection. (<b>a</b>) Visualization of original ATL03 footprint data: The horizontal axis represents the distance along the orbital direction, and the vertical axis represents the altitude of the footprints; (<b>b</b>) Distribution of the terrain trend line: The horizontal coordinate of an image pixel is represented by X, while the vertical coordinate is represented by Y.</p>
Full article ">Figure 8
<p>The accuracy assessment of terrain trend line extraction.</p>
Full article ">Figure 9
<p>The illustration of the signal footprints recognition. The horizontal axis represents the distance along the orbital direction, and the vertical axis represents the altitude of the footprints. The blue data points represent the airborne laser scanner (ALS) digital elevation model (DEM) results, while the green data points represent the ALS digital surface model (DSM) results. (<b>a</b>) The rough denoising result of ATL03; (<b>b</b>) The accurate denoising result of ATL03.</p>
Full article ">Figure 10
<p>The accuracy assessment of extracted forest structural parameters through independent sample testing. (<b>a</b>–<b>c</b>) Accuracy of forest height using CNN, ResNet50, and EfficientNetB3; (<b>d</b>–<b>f</b>) Accuracy of canopy cover using CNN, ResNet50, and EfficientNetB3.</p>
Full article ">Figure 11
<p>The influence of parameters on the accuracy of the CNN forest height estimation model. (<b>a</b>) The influence of learning rate on accuracy; (<b>b</b>) The influence of iteration on accuracy; (<b>c</b>) The influence of batch size on accuracy; (<b>d</b>) The influence of image size on accuracy.</p>
Full article ">
19 pages, 1688 KiB  
Article
Machine Learning-Based Control of Autonomous Vehicles for Solar Panel Cleaning Systems in Agricultural Solar Farms
by Farima Hajiahmadi, Mohammad Jafari and Mahmut Reyhanoglu
AgriEngineering 2024, 6(2), 1417-1435; https://doi.org/10.3390/agriengineering6020081 - 20 May 2024
Viewed by 1246
Abstract
This paper presents a machine learning (ML)-based approach for the intelligent control of Autonomous Vehicles (AVs) utilized in solar panel cleaning systems, aiming to mitigate challenges arising from uncertainties, disturbances, and dynamic environments. Solar panels, predominantly situated in dedicated lands for solar energy [...] Read more.
This paper presents a machine learning (ML)-based approach for the intelligent control of Autonomous Vehicles (AVs) utilized in solar panel cleaning systems, aiming to mitigate challenges arising from uncertainties, disturbances, and dynamic environments. Solar panels, predominantly situated in dedicated lands for solar energy production (e.g., agricultural solar farms), are susceptible to dust and debris accumulation, leading to diminished energy absorption. Instead of labor-intensive manual cleaning, robotic cleaners offer a viable solution. AVs equipped to transport and precisely position these cleaning robots are indispensable for the efficient navigation among solar panel arrays. However, environmental obstacles (e.g., rough terrain), variations in solar panel installation (e.g., height disparities, different angles), and uncertainties (e.g., AV and environmental modeling) may degrade the performance of traditional controllers. In this study, a biologically inspired method based on Brain Emotional Learning (BEL) is developed to tackle the aforementioned challenges. The developed controller is implemented numerically using MATLAB-SIMULINK. The paper concludes with a comparative analysis of the AVs’ performance using both PID and developed controllers across various scenarios, highlighting the efficacy and advantages of the intelligent control approach for AVs deployed in solar panel cleaning systems within agricultural solar farms. Simulation results demonstrate the superior performance of the ML-based controller, showcasing significant improvements over the PID controller. Full article
Show Figures

Figure 1

Figure 1
<p>Autonomous Vehicle utilized for carrying and positioning cleaning robots in solar panel cleaning systems.</p>
Full article ">Figure 2
<p>Graphical schematic depicting the mathematical model of Brain Emotional Learning (BEL) in the mammalian limbic system.</p>
Full article ">Figure 3
<p>Brain Emotional Learning controller architecture for closed-loop control of Autonomous Vehicles.</p>
Full article ">Figure 4
<p>The desired and actual outputs of the system [the displacement (<span class="html-italic">x</span>) and the angles of the front and rear scissors (<math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math>)] in constant trajectory tracking (see Scenario I: Maintaining constant trajectories for AV angles and displacement). The developed BEL-based controller is in red (dashed line), the PID is in blue (dashed–dotted line), and the desired trajectories are in green (solid line).</p>
Full article ">Figure 5
<p>The generated forces and torques by both controllers in constant trajectory tracking (see Scenario I: Maintaining constant trajectories for AV angles and displacement). The developed BEL-based controller is in magenta (dashed line), and the PID is in cyan (dashed–dotted line).</p>
Full article ">Figure 6
<p>The tracking errors in constant trajectory tracking (see Scenario I: Maintaining constant trajectories for AV angles and displacement). The developed BEL-based controller is in orange (dashed line), and the PID is in purple (dashed–dotted line).</p>
Full article ">Figure 7
<p>The desired and actual outputs of the system [the displacement (<span class="html-italic">x</span>) and the angles of the front and rear scissors (<math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math>)] in tracking constant/time-varying sinusoidal trajectories (see Scenario II: Tracking sinusoidal trajectories to maintain AV angles). The developed BEL-based controller is in red (dashed line), the PID is in blue (dashed–dotted line), and the desired trajectories are in green (solid line).</p>
Full article ">Figure 8
<p>The generated forces and torques by both controllers in tracking constant/time-varying sinusoidal trajectories (see Scenario II: Tracking sinusoidal trajectories to maintain AV angles). The developed BEL-based controller is in magenta (dashed line) and the PID is in cyan (dashed–dotted line).</p>
Full article ">Figure 9
<p>The tracking errors in tracking constant/time-varying sinusoidal trajectories (see Scenario II: Tracking sinusoidal trajectories to maintain AV angles). The developed BEL-based controller is in orange (dashed line), and the PID is in purple (dashed–dotted line).</p>
Full article ">Figure 10
<p>The desired and actual outputs of the system [the displacement (<span class="html-italic">x</span>) and the angles of the front and rear scissors (<math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math>)] in tracking trajectories in the presence of substantial external disturbances (see Scenario III: Preserving AV angles and displacement in the presence of substantial external disturbances). External disturbances are introduced to perturb the angles of the front and rear scissors within the time intervals of [8–9] and [14–15] seconds, respectively. The developed BEL-based controller is in red (dashed line), the PID is in blue (dashed–dotted line), and the desired trajectories are in green (solid line).</p>
Full article ">Figure 11
<p>The generated forces and torques by both controllers in tracking trajectories in the presence of substantial external disturbances (see Scenario III: Preserving AV angles and displacement in the presence of substantial external disturbances). The developed BEL-based controller is in magenta (dashed line), and the PID is in cyan (dashed–dotted line).</p>
Full article ">Figure 12
<p>The tracking errors in tracking trajectories in the presence of substantial external disturbances (see Scenario III: Preserving AV angles and displacement in the presence of substantial external disturbances). The developed BEL-based controller is in orange (dashed line), and the PID is in purple (dashed–dotted line).</p>
Full article ">
18 pages, 3875 KiB  
Article
Estimation of Soil Characteristic Parameters for Electric Mountain Tractor Based on Gauss–Newton Iteration Method
by Zhiqiang Xi, Tian Feng, Zhijun Liu, Huaijun Xu, Jingyang Zheng and Liyou Xu
World Electr. Veh. J. 2024, 15(5), 217; https://doi.org/10.3390/wevj15050217 - 15 May 2024
Cited by 1 | Viewed by 819
Abstract
Future field work tasks will require mountain tractors to pass through rough terrain with limited human supervision. The wheel–soil interaction plays a critical role in rugged terrain mobility. In this paper, an algorithm for the estimation of soil characteristic parameters based on the [...] Read more.
Future field work tasks will require mountain tractors to pass through rough terrain with limited human supervision. The wheel–soil interaction plays a critical role in rugged terrain mobility. In this paper, an algorithm for the estimation of soil characteristic parameters based on the Simpson numerical integration method and Gauss–Newton iteration method is presented. These parameters can be used for passability prediction or in a traction control algorithm to improve tractor mobility and to plan safe operation paths for autonomous navigation systems. To verify the effectiveness of the solving algorithm, different initial values and soils were selected for simulation calculations of soil characteristic parameters such as internal friction angle, settlement index, and the joint parameter of soil cohesion modulus and friction modulus. The results show that the error was kept within 2%, and the calculation time did not exceed 0.84 s, demonstrating high robustness and real-time performance. To test the applicability of the algorithm model, further research was conducted using different wheel parameters of electric mountain tractors under wet clay conditions. The results show that these parameters also have high accuracy and stability with only a few iterations. Thus, the estimation algorithm can meet the requirements of quickly and accurately identifying soil characteristic parameters during tractor operation. A criterion for the passability of wheeled tractors through unknown terrain is proposed, utilizing identified soil parameters. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of force distribution interaction between rigid wheel and soft ground.</p>
Full article ">Figure 2
<p>Comparison of the distribution curves of the improved shear stress model and the original shear model.</p>
Full article ">Figure 3
<p>Vector diagram of the velocity at point P on the outer edge of the wheel.</p>
Full article ">Figure 4
<p>Algorithm framework for estimation of soil characteristic parameters.</p>
Full article ">Figure 5
<p>Block diagram of Gauss–Newton iterative solution method.</p>
Full article ">Figure 6
<p>Iteration of soil internal friction angle for different soils based on the condition of 20% initial value error from the true value. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 7
<p>Iteration of joint parameter for different soils based on the condition of 20% initial value error from the true value condition. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 8
<p>Iteration of subsidence index for different soils based on the condition of 20% initial value error from the true value. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 9
<p>Iteration of soil internal friction angle for different soils based on the condition of 50% initial value error from the true value. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 10
<p>Iteration of the joint parameter for different soils based on the condition of 50% initial value error from the true value. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 11
<p>Iteration of the subsidence index for different soils based on the condition of 50% initial value error from the true value. (<b>a</b>) Based on wet clay; (<b>b</b>) based on sandy loam.</p>
Full article ">Figure 12
<p>Traversability criterion for wheeled tractors.</p>
Full article ">Figure 13
<p>Drive torques required for wheel tractor to traverse two distinct terrains.</p>
Full article ">
Back to TopTop