[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (115)

Search Parameters:
Keywords = visual-inertial odometry

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 2929 KiB  
Article
R-LVIO: Resilient LiDAR-Visual-Inertial Odometry for UAVs in GNSS-denied Environment
by Bing Zhang, Xiangyu Shao, Yankun Wang, Guanghui Sun and Weiran Yao
Drones 2024, 8(9), 487; https://doi.org/10.3390/drones8090487 (registering DOI) - 14 Sep 2024
Viewed by 194
Abstract
In low-altitude, GNSS-denied scenarios, Unmanned aerial vehicles (UAVs) rely on sensor fusion for self-localization. This article presents a resilient multi-sensor fusion localization system that integrates light detection and ranging (LiDAR), cameras, and inertial measurement units (IMUs) to achieve state estimation for UAVs. To [...] Read more.
In low-altitude, GNSS-denied scenarios, Unmanned aerial vehicles (UAVs) rely on sensor fusion for self-localization. This article presents a resilient multi-sensor fusion localization system that integrates light detection and ranging (LiDAR), cameras, and inertial measurement units (IMUs) to achieve state estimation for UAVs. To address challenging environments, especially unstructured ones, IMU predictions are used to compensate for pose estimation in the visual and LiDAR components. Specifically, the accuracy of IMU predictions is enhanced by increasing the correction frequency of IMU bias through data integration from the LiDAR and visual modules. To reduce the impact of random errors and measurement noise in LiDAR points on visual depth measurement, cross-validation of visual feature depth is performed using reprojection error to eliminate outliers. Additionally, a structure monitor is introduced to switch operation modes in hybrid point cloud registration, ensuring accurate state estimation in both structured and unstructured environments. In unstructured scenes, a geometric primitive capable of representing irregular planes is employed for point-to-surface registration, along with a novel pose-solving method to estimate the UAV’s pose. Both private and public datasets collected by UAVs validate the proposed system, proving that it outperforms state-of-the-art algorithms by at least 12.6%. Full article
25 pages, 4182 KiB  
Article
W-VSLAM: A Visual Mapping Algorithm for Indoor Inspection Robots
by Dingji Luo, Yucan Huang, Xuchao Huang, Mingda Miao and Xueshan Gao
Sensors 2024, 24(17), 5662; https://doi.org/10.3390/s24175662 - 30 Aug 2024
Viewed by 416
Abstract
In recent years, with the widespread application of indoor inspection robots, high-precision, robust environmental perception has become essential for robotic mapping. Addressing the issues of visual–inertial estimation inaccuracies due to redundant pose degrees of freedom and accelerometer drift during the planar motion of [...] Read more.
In recent years, with the widespread application of indoor inspection robots, high-precision, robust environmental perception has become essential for robotic mapping. Addressing the issues of visual–inertial estimation inaccuracies due to redundant pose degrees of freedom and accelerometer drift during the planar motion of mobile robots in indoor environments, we propose a visual SLAM perception method that integrates wheel odometry information. First, the robot’s body pose is parameterized in SE(2) and the corresponding camera pose is parameterized in SE(3). On this basis, we derive the visual constraint residuals and their Jacobian matrices for reprojection observations using the camera projection model. We employ the concept of pre-integration to derive pose-constraint residuals and their Jacobian matrices and utilize marginalization theory to derive the relative pose residuals and their Jacobians for loop closure constraints. This approach solves the nonlinear optimization problem to obtain the optimal pose and landmark points of the ground-moving robot. A comparison with the ORBSLAM3 algorithm reveals that, in the recorded indoor environment datasets, the proposed algorithm demonstrates significantly higher perception accuracy, with root mean square error (RMSE) improvements of 89.2% in translation and 98.5% in rotation for absolute trajectory error (ATE). The overall trajectory localization accuracy ranges between 5 and 17 cm, validating the effectiveness of the proposed algorithm. These findings can be applied to preliminary mapping for the autonomous navigation of indoor mobile robots and serve as a basis for path planning based on the mapping results. Full article
Show Figures

Figure 1

Figure 1
<p>Block diagram of mobile robot visual SLAM with integrated wheel speed.</p>
Full article ">Figure 2
<p>Schematic diagram of mobile robot coordinate system.</p>
Full article ">Figure 3
<p>Camera projection model.</p>
Full article ">Figure 4
<p>Schematic diagram of wheel speed information pre-integration.</p>
Full article ">Figure 5
<p>Schematic diagram of square movement in the comparative experiment.</p>
Full article ">Figure 6
<p>Reference keyframe and current frame in previous tracking, (<b>a</b>) reference keyframe of previous tracking and (<b>b</b>) current frame of previous tracking.</p>
Full article ">Figure 7
<p>Comparative Experiment One: robot poses and environmental map points obtained by W-VSLAM.</p>
Full article ">Figure 8
<p>Comparative Experiment Two: robot poses and environmental map points obtained by W-VSLAM.</p>
Full article ">Figure 9
<p>Comparative Experiment One: trajectory comparison chart of different algorithms.</p>
Full article ">Figure 10
<p>Comparative Experiment Two: trajectory comparison chart of different algorithms.</p>
Full article ">Figure 11
<p>Comparative test one: translational component comparison chart of trajectories from different algorithms.</p>
Full article ">Figure 12
<p>(<b>a</b>) Perception results of robot poses and map points in Experiment One. (<b>b</b>) Comparison between Experiment One and reference trajectory.</p>
Full article ">Figure 13
<p>(<b>a</b>) Perception results of robot poses and map points in Experiment Two. (<b>b</b>) Comparison between Experiment Two and reference trajectory.</p>
Full article ">Figure 14
<p>(<b>a</b>) Perception results of robot poses and map points in Experiment Three. (<b>b</b>) Comparison between Experiment Three and reference trajectory.</p>
Full article ">Figure 15
<p>Indoor long corridor environment trajectory; Rivz result diagram.</p>
Full article ">Figure 16
<p>Comparison chart of trajectories in the indoor long corridor environment.</p>
Full article ">Figure 17
<p>Comparison chart of translational components of trajectories in the indoor long corridor environment.</p>
Full article ">Figure 18
<p>Absolute accuracy of rotational estimation in the indoor long corridor environment.</p>
Full article ">Figure 19
<p>Relative accuracy of rotational estimation in the indoor long corridor environment (with a 1° increment).</p>
Full article ">
14 pages, 3371 KiB  
Technical Note
Pose Estimation Based on Bidirectional Visual–Inertial Odometry with 3D LiDAR (BV-LIO)
by Gang Peng, Qiang Gao, Yue Xu, Jianfeng Li, Zhang Deng and Cong Li
Remote Sens. 2024, 16(16), 2970; https://doi.org/10.3390/rs16162970 - 14 Aug 2024
Viewed by 653
Abstract
Due to the limitation of a single sensor such as only camera or only LiDAR, the Visual SLAM detects few effective features in the case of poor lighting or no texture. The LiDAR SLAM will also degrade in an unstructured environment and open [...] Read more.
Due to the limitation of a single sensor such as only camera or only LiDAR, the Visual SLAM detects few effective features in the case of poor lighting or no texture. The LiDAR SLAM will also degrade in an unstructured environment and open spaces, which reduces the accuracy of pose estimation and the quality of mapping. In order to solve this problem, on account of the high efficiency of Visual odometry and the high accuracy of LiDAR odometry, this paper investigates the multi-sensor fusion of bidirectional visual–inertial odometry with 3D LiDAR for pose estimation. This method can couple the IMU with the bidirectional vision respectively, and the LiDAR odometry is obtained assisted by the bidirectional visual inertial. The factor graph optimization is constructed, which effectively improves the accuracy of pose estimation. The algorithm in this paper is compared with LIO-LOAM, LeGO-LOAM, VINS-Mono, and so on using challenging datasets such as KITTI and M2DGR. The results show that this method effectively improves the accuracy of pose estimation and has high application value for mobile robots. Full article
Show Figures

Figure 1

Figure 1
<p>The System Overview of BV-LIO. The <b>blue</b> sections represent input modules (LiDAR, IMU, and camera inputs), the <b>orange</b> sections are LiDAR processing and odometry modules (LiDAR data feature extraction, odometry, and initial depth association), the red sections indicate camera processing and odometry modules (bidirectional vision inertial fusion and factor graph optimization), the gray sections show loop closure detection (vision and LiDAR loop closure detection), and the yellow section is the output module (final pose estimation determination).</p>
Full article ">Figure 2
<p>The maintenance time change diagram of local point cloud set.</p>
Full article ">Figure 3
<p>The loop closure detection results: (<b>a</b>) front camera. (<b>b</b>) back camera. The loop closure detection successfully identified 54 pairs of corresponding points from the front camera and 89 pairs from the back camera.</p>
Full article ">Figure 4
<p>Visual SLAM algorithm degradation in dark night scene (M2DGR_04).</p>
Full article ">Figure 5
<p>The EVO Trajectory drawing (<b>a</b>) <b>KITTI_05</b> (day, long term, small loop back) (<b>b</b>) <b>KITTI_08</b> (day, long term, small loop back) (<b>c</b>) <b>KITTI_10</b> (day, one turn) (<b>d</b>) <b>M2DGR_2</b> (day, long term) (<b>e</b>) <b>M2DGR_04</b> (night, around lawn, loop back) (<b>f</b>) <b>M2DGR_06</b> (night, one turn).</p>
Full article ">Figure 6
<p>(<b>a</b>) The Diagram of Mobile Robot and Sensor Suite. (<b>b</b>) Corridor Diagram of Inspection Scene.</p>
Full article ">Figure 7
<p>Trajectory Estimated in Corridor Scene.</p>
Full article ">
25 pages, 7789 KiB  
Article
Mix-VIO: A Visual Inertial Odometry Based on a Hybrid Tracking Strategy
by Huayu Yuan, Ke Han and Boyang Lou
Sensors 2024, 24(16), 5218; https://doi.org/10.3390/s24165218 - 12 Aug 2024
Viewed by 747
Abstract
In this paper, we proposed Mix-VIO, a monocular and binocular visual-inertial odometry, to address the issue where conventional visual front-end tracking often fails under dynamic lighting and image blur conditions. Mix-VIO adopts a hybrid tracking approach, combining traditional handcrafted tracking techniques with Deep [...] Read more.
In this paper, we proposed Mix-VIO, a monocular and binocular visual-inertial odometry, to address the issue where conventional visual front-end tracking often fails under dynamic lighting and image blur conditions. Mix-VIO adopts a hybrid tracking approach, combining traditional handcrafted tracking techniques with Deep Neural Network (DNN)-based feature extraction and matching pipelines. The system employs deep learning methods for rapid feature point detection, while integrating traditional optical flow methods and deep learning-based sparse feature matching methods to enhance front-end tracking performance under rapid camera motion and environmental illumination changes. In the back-end, we utilize sliding window and bundle adjustment (BA) techniques for local map optimization and pose estimation. We conduct extensive experimental validations of the hybrid feature extraction and matching methods, demonstrating the system’s capability to maintain optimal tracking results under illumination changes and image blur. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>b</b>) show adjacent images from the EuRoc dataset [<a href="#B12-sensors-24-05218" class="html-bibr">12</a>], where noticeable blurring occurs between images. (<b>c</b>,<b>d</b>) display adjacent images from the UMA-VI dataset [<a href="#B13-sensors-24-05218" class="html-bibr">13</a>], highlighting significant changes in lighting between the two images. Blue indicates fewer tracking instances, representing initial feature point, while red indicates frequent successful feature matches due to multiple trackings. Points with rings illustrate results obtained through traditional feature extraction and optical flow matching; the inner circle radius represents the suppression radius for SP features, and the outer circle radius for traditional features, thus dispersing the feature points. The green arrows point to the positions of the points in the previous frame from current frame. Points without rings represent SP features, with 1024 features extracted in the image. It can be observed that in (<b>a</b>,<b>b</b>), despite the blurring, optical flow matching still matches many feature points, but SP + LG largely fails. In (<b>c</b>,<b>d</b>), due to drastic illumination changes, optical flow matching fails, but SP + LG still successfully matches many feature points. The traditional approach and deep learning approach complement each other, achieving better tracking results.</p>
Full article ">Figure 2
<p>Mix-VIO system overview.</p>
Full article ">Figure 3
<p>Mixed-up feature-tracking pipeline overview.</p>
Full article ">Figure 4
<p>Superpoint network architecture.</p>
Full article ">Figure 5
<p>Lightglue network using Superpoint as the input.</p>
Full article ">Figure 6
<p>The process of point tracking and the hybrid feature-tracking strategy. The red-colored points represent points where optical flow tracking fails; green points represent Superpoint features which are matched using Lightglue; blue points represent points tracked by optical flow. It is noteworthy that yellow Superpoint points, although unmatched, are added to the system if the number of feature points does not reach the threshold, serving as a basis for feature matching in the next frame.</p>
Full article ">Figure 7
<p>Sequences (Machine House, MH) and collection equipment in EuRoc dataset.</p>
Full article ">Figure 8
<p>Comparison of SP + LG (<b>a</b>,<b>b</b>) and GFT+optical flow (<b>c</b>,<b>d</b>) methods under image blur caused by fast camera movement speed. To distinguish between the two, connect the points matched by the optical flow method with lines and represent the SP points with hollow circles.</p>
Full article ">Figure 9
<p>Comparison of another set of SP+LG (<b>a</b>,<b>b</b>) and GFT + optical flow (<b>c</b>,<b>d</b>) methods under image blur caused by rapid camera movement.</p>
Full article ">Figure 10
<p>Sequences (ill-change) in UMA-VI dataset. We selected some representative images, and in the actual sequence, several images are kept in low-light conditions. From left to right, the lighting in each row gradually dims and then is turned back on.</p>
Full article ">Figure 11
<p>Comparison of another set of SP + LG (<b>a</b>,<b>b</b>) and GFT + optical flow (<b>c</b>,<b>d</b>) methods under illumination variations caused by the lighting change. Even if completely dark images are skipped, the optical flow method still cannot track the results of the previous and subsequent frames, which will cause the VIO system to fail.</p>
Full article ">Figure 12
<p>Comparison of another set of SP + LG (<b>a</b>,<b>b</b>) and GFT + optical flow (<b>c</b>,<b>d</b>) methods under illumination variation caused by the lighting change in the UMA-VI dataset. Although it is no longer possible to clearly distinguish the contour, the SP + LG-based method can still achieve good tracking results.</p>
Full article ">
22 pages, 5331 KiB  
Article
Rapid Initialization Method of Unmanned Aerial Vehicle Swarm Based on VIO-UWB in Satellite Denial Environment
by Runmin Wang and Zhongliang Deng
Drones 2024, 8(7), 339; https://doi.org/10.3390/drones8070339 - 22 Jul 2024
Viewed by 586
Abstract
In environments where satellite signals are blocked, initializing UAV swarms quickly is a technical challenge, especially indoors or in areas with weak satellite signals, making it difficult to establish the relative position of the swarm. Two common methods for initialization are using the [...] Read more.
In environments where satellite signals are blocked, initializing UAV swarms quickly is a technical challenge, especially indoors or in areas with weak satellite signals, making it difficult to establish the relative position of the swarm. Two common methods for initialization are using the camera for joint SLAM initialization, which increases communication burden due to image feature point analysis, and obtaining a rough positional relationship using prior information through a device such as a magnetic compass, which lacks accuracy. In recent years, visual–inertial odometry (VIO) technology has significantly progressed, providing new solutions. With improved computing power and enhanced VIO accuracy, it is now possible to establish the relative position relationship through the movement of drones. This paper proposes a two-stage robust initialization method for swarms of more than four UAVs, suitable for larger-scale satellite denial scenarios. Firstly, the paper analyzes the Cramér–Rao lower bound (CRLB) problem and the moving configuration problem of the cluster to determine the optimal anchor node for the algorithm. Subsequently, a strategy is used to screen anchor nodes that are close to the lower bound of CRLB, and an optimization problem is constructed to solve the position relationship between anchor nodes through the relative motion and ranging relationship between UAVs. This optimization problem includes quadratic constraints as well as linear constraints and is a quadratically constrained quadratic programming problem (QCQP) with high robustness and high precision. After addressing the anchor node problem, this paper simplifies and improves a fast swarm cooperative positioning algorithm, which is faster than the traditional multidimensional scaling (MDS) algorithm. The results of theoretical simulations and actual UAV tests demonstrate that the proposed algorithm is advanced, superior, and effectively solves the UAV swarm initialization problem under the condition of a satellite signal rejection. Full article
Show Figures

Figure 1

Figure 1
<p>Model of drone swarm.</p>
Full article ">Figure 2
<p>FIM (Fisher information matrix) simulation analysis.</p>
Full article ">Figure 3
<p>D<sup>2</sup> algorithm simulation analysis.</p>
Full article ">Figure 4
<p>The simulation results of the improved SMACOF algorithm compared with the DC-MDS algorithm.</p>
Full article ">Figure 5
<p>Algorithm flowchart of this paper.</p>
Full article ">Figure 6
<p>A model of solving the anchor problem.</p>
Full article ">Figure 7
<p>The different trajectories used in the simulation experiment.</p>
Full article ">Figure 8
<p>Simulation results. The solid line in the figure represents the influence of different numbers of drones on the convergence time of the algorithm, and the dotted line represents the impact of different numbers of drones on the accuracy of the algorithm.</p>
Full article ">Figure 9
<p>(<b>a</b>) The quadrotor platforms used in our experiments; (<b>b</b>) the flight site used in our experiment.</p>
Full article ">Figure 10
<p>The yaw angle of the anchor node is the result after time alignment. The yellow line represents the true value, and the green line represents the result of the VIO output.</p>
Full article ">Figure 11
<p>The trajectories of the anchor node are the result after time alignment. The red line represents the true value, and the blue line represents the result of the VIO output.</p>
Full article ">Figure 12
<p>Solver time of different algorithms. The green boxes represent nonlinear least squares (NLS); the purple box represents semidefinite programming (SDP); the light blue box represents the quadratic convex optimization problem (QCQP); and the red box represents the method proposed in this paper.</p>
Full article ">Figure 13
<p>Estimation results in our experiment.</p>
Full article ">
19 pages, 3876 KiB  
Article
An Adaptive Fast Incremental Smoothing Approach to INS/GPS/VO Factor Graph Inference
by Zhaoxu Tian, Yongmei Cheng and Shun Yao
Appl. Sci. 2024, 14(13), 5691; https://doi.org/10.3390/app14135691 - 29 Jun 2024
Viewed by 482
Abstract
In response to asynchronous and delayed sensors within multi-sensor integrated navigation systems, the computational complexity of joint optimization navigation solutions persistently rises. This paper introduces an adaptive fast integrated navigation algorithm for INS/GPS/VO based on factor graph. The factor graph model for INS/GPS/VO [...] Read more.
In response to asynchronous and delayed sensors within multi-sensor integrated navigation systems, the computational complexity of joint optimization navigation solutions persistently rises. This paper introduces an adaptive fast integrated navigation algorithm for INS/GPS/VO based on factor graph. The factor graph model for INS/GPS/VO is developed subsequent to individual modeling of the Inertial Navigation System (INS), Global Positioning System (GPS), and Visual Odometer (VO) using the factor graph model approach. Additionally, an Adaptive Fast Incremental Smoothing (AFIS) factor graph optimization algorithm is proposed. The simulation results demonstrate that the factor-graph-based integrated navigation algorithm consistently yields high-precision navigation outcomes even amidst dynamic changes in sensor validity and the presence of asynchronous and delayed sensor measurements. Notably, the AFIS factor graph optimization algorithm significantly enhances real-time performance compared to traditional Incremental Smoothing (IF) algorithms, while maintaining comparable real-time accuracy. Full article
(This article belongs to the Collection Advances in Automation and Robotics)
Show Figures

Figure 1

Figure 1
<p>Construct scheme of INS factor graph model.</p>
Full article ">Figure 2
<p>The comparison between IMU factor and equivalent IMU factor. (<b>a</b>) Factor graph model using IMU factor; (<b>b</b>) Factor graph model using equivalent IMU factor.</p>
Full article ">Figure 3
<p>Factor graph model including factors and nodes at different times. (<b>a</b>,<b>b</b>) Factor graphs that include GPS, prior, equivalent IMU, and bias factors at time <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>2</mn> </msub> </mrow> </semantics></math>; (<b>c</b>) Factor graphs with VO, prior, equivalent IMU, and bias factors; (<b>d</b>) Factor graphs with GPS, VO, prior, IMU, and bias factors. Which is equivalent to the factor graph represented in (<b>c</b>), utilizing conventional IMU factors. Assuming that GPS and VO measurements are acquired every 50 IMU measurements. Additionally, bias nodes are incorporated at the same frequency as depicted in (<b>c</b>).</p>
Full article ">Figure 4
<p>Factor graph model for asynchronous, delayed, and missing measurements. (<b>a</b>) GPS and VO measurement frequency are the same as the system output frequency; (<b>b</b>) GPS and VO measurement frequency lower than system output frequency, and the missing GPS measurement (measurement <math display="inline"><semantics> <mrow> <msubsup> <mi>f</mi> <mn>2</mn> <mrow> <mi>G</mi> <mi>P</mi> <mi>S</mi> </mrow> </msubsup> </mrow> </semantics></math> is missing).</p>
Full article ">Figure 5
<p>Bayesian network corresponding to the factor graph at different times. (<b>a</b>) Bayesian network corresponding to the elimination of the factor graph at time <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>1</mn> </msub> </mrow> </semantics></math>; (<b>b</b>) Incorporating new measurements into a factor graph at time <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mn>2</mn> </msub> </mrow> </semantics></math>, introducing additional nodes as <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>2</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>2</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Schematic diagram of IS (the changed variables are shown in red).</p>
Full article ">Figure 7
<p>Schematic diagram of AFIS (the changed variables are shown in red).</p>
Full article ">Figure 8
<p>Program flow chart of the AFIS algorithm.</p>
Full article ">Figure 9
<p>Position error of KF and FG with dynamic changes in GPS effectiveness.</p>
Full article ">Figure 10
<p>Position error of KF and FG in different measurement delays.</p>
Full article ">Figure 11
<p>Radial position error of different algorithms in different scenarios. (<b>a</b>) Position error in scenario 1; (<b>b</b>) Position error in scenario 2.</p>
Full article ">Figure 12
<p>Comparison of total runtime of different algorithms in different scenarios. (<b>a</b>) Total runtime in scenario 1; (<b>b</b>) Total runtime in scenario 2.</p>
Full article ">
22 pages, 4067 KiB  
Article
A Sensor Fusion Approach to Observe Quadrotor Velocity
by José Ramón Meza-Ibarra, Joaquín Martínez-Ulloa, Luis Alfonso Moreno-Pacheco and Hugo Rodríguez-Cortés
Sensors 2024, 24(11), 3605; https://doi.org/10.3390/s24113605 - 3 Jun 2024
Viewed by 612
Abstract
The growing use of Unmanned Aerial Vehicles (UAVs) raises the need to improve their autonomous navigation capabilities. Visual odometry allows for dispensing positioning systems, such as GPS, especially on indoor flights. This paper reports an effort toward UAV autonomous navigation by proposing a [...] Read more.
The growing use of Unmanned Aerial Vehicles (UAVs) raises the need to improve their autonomous navigation capabilities. Visual odometry allows for dispensing positioning systems, such as GPS, especially on indoor flights. This paper reports an effort toward UAV autonomous navigation by proposing a translational velocity observer based on inertial and visual measurements for a quadrotor. The proposed observer complementarily fuses available measurements from different domains and is synthesized following the Immersion and Invariance observer design technique. A formal Lyapunov-based observer error convergence to zero is provided. The proposed observer algorithm is evaluated using numerical simulations in the Parrot Mambo Minidrone App from Simulink-Matlab. Full article
(This article belongs to the Collection Navigation Systems and Sensors)
Show Figures

Figure 1

Figure 1
<p>Mambo Parrot’s coordinate frames.</p>
Full article ">Figure 2
<p>Pinhole camera principle [<a href="#B30-sensors-24-03605" class="html-bibr">30</a>].</p>
Full article ">Figure 3
<p>Sensor fusion.</p>
Full article ">Figure 4
<p>Measured acceleration <math display="inline"><semantics> <msubsup> <mi>a</mi> <mi>x</mi> <mi>b</mi> </msubsup> </semantics></math> and reconstructed acceleration <math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>a</mi> <mo>¯</mo> </mover> <mi>x</mi> <mi>b</mi> </msubsup> </semantics></math>.</p>
Full article ">Figure 5
<p>Optical flow block design.</p>
Full article ">Figure 6
<p>Aircraft tracking: circular-type trajectory.</p>
Full article ">Figure 7
<p>Aircraft tracking: square-type trajectory.</p>
Full article ">Figure 8
<p>Specific force measured along the <math display="inline"><semantics> <mrow> <mn>0</mn> <msup> <mi>X</mi> <mi>b</mi> </msup> </mrow> </semantics></math> axis while the quadrotor follows a circular trajectory.</p>
Full article ">Figure 9
<p>Specific force measured along the <math display="inline"><semantics> <mrow> <mn>0</mn> <msup> <mi>X</mi> <mi>b</mi> </msup> </mrow> </semantics></math> axis while the quadrotor follows a square-type trajectory.</p>
Full article ">Figure 10
<p>Computed optical flow along the <math display="inline"><semantics> <mrow> <mn>0</mn> <msup> <mi>X</mi> <mi>b</mi> </msup> </mrow> </semantics></math> axis while the quadrotor tracks a circular trajectory.</p>
Full article ">Figure 11
<p>Computed optical flow along the <math display="inline"><semantics> <mrow> <mn>0</mn> <msup> <mi>X</mi> <mi>b</mi> </msup> </mrow> </semantics></math> axis while the quadrotor tracks a square-type trajectory.</p>
Full article ">Figure 12
<p>Observed speed <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>u</mi> <mo>^</mo> </mover> <mo>+</mo> <msub> <mo>Γ</mo> <mn>11</mn> </msub> <msub> <mi>σ</mi> <mn>1</mn> </msub> </mrow> </semantics></math> (blue line) and speed computed by the Parrot Mambo simulator algorithm <span class="html-italic">u</span> (yellow line).</p>
Full article ">Figure 13
<p>Observed speed <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> <mo>+</mo> <msub> <mo>Γ</mo> <mn>22</mn> </msub> <msub> <mi>σ</mi> <mn>2</mn> </msub> </mrow> </semantics></math> (blue line) and speed computed by the Parrot Mambo simulator algorithm <span class="html-italic">v</span> (yellow line).</p>
Full article ">Figure 14
<p>Observed speed <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>u</mi> <mo>^</mo> </mover> <mo>+</mo> <msub> <mo>Γ</mo> <mn>11</mn> </msub> <msub> <mi>σ</mi> <mn>1</mn> </msub> </mrow> </semantics></math> (blue line) and speed computed by the Parrot Mambo simulator algorithm <span class="html-italic">u</span> (yellow line).</p>
Full article ">Figure 15
<p>Observed speed <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> <mo>+</mo> <msub> <mo>Γ</mo> <mn>22</mn> </msub> <msub> <mi>σ</mi> <mn>2</mn> </msub> </mrow> </semantics></math> (blue line) and speed computed by the Parrot Mambo simulator algorithm <span class="html-italic">v</span> (yellow line).</p>
Full article ">Figure 16
<p>Observed speeds errors for the circular trajectory. <math display="inline"><semantics> <mover accent="true"> <mi>u</mi> <mo>˜</mo> </mover> </semantics></math> (blue line), <math display="inline"><semantics> <mover accent="true"> <mi>v</mi> <mo>˜</mo> </mover> </semantics></math> (red line).</p>
Full article ">Figure 17
<p>Observed speeds errors for the square-type trajectory. <math display="inline"><semantics> <mover accent="true"> <mi>u</mi> <mo>˜</mo> </mover> </semantics></math> (blue line), <math display="inline"><semantics> <mover accent="true"> <mi>v</mi> <mo>˜</mo> </mover> </semantics></math> (red line).</p>
Full article ">Figure 18
<p>Eigenvalues for different combinations of gains <math display="inline"><semantics> <msub> <mi>α</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>α</mi> <mn>2</mn> </msub> </semantics></math>.</p>
Full article ">Figure 19
<p>Eigenvalues vs. observer’s estimation error in the <math display="inline"><semantics> <mrow> <mn>0</mn> <msup> <mi>X</mi> <mi>b</mi> </msup> </mrow> </semantics></math> axis.</p>
Full article ">
26 pages, 18657 KiB  
Article
Development of Unmanned Aerial Vehicle Navigation and Warehouse Inventory System Based on Reinforcement Learning
by Huei-Yung Lin, Kai-Lun Chang and Hsin-Ying Huang
Drones 2024, 8(6), 220; https://doi.org/10.3390/drones8060220 - 28 May 2024
Cited by 1 | Viewed by 1031
Abstract
In this paper, we present the exploration of indoor positioning technologies for UAVs, as well as navigation techniques for path planning and obstacle avoidance. The objective was to perform warehouse inventory tasks, using a drone to search for barcodes or markers to identify [...] Read more.
In this paper, we present the exploration of indoor positioning technologies for UAVs, as well as navigation techniques for path planning and obstacle avoidance. The objective was to perform warehouse inventory tasks, using a drone to search for barcodes or markers to identify objects. For the indoor positioning techniques, we employed visual-inertial odometry (VIO), ultra-wideband (UWB), AprilTag fiducial markers, and simultaneous localization and mapping (SLAM). These algorithms included global positioning, local positioning, and pre-mapping positioning, comparing the merits and drawbacks of various techniques and trajectories. For UAV navigation, we combined the SLAM-based RTAB-map indoor mapping and navigation path planning of the ROS for indoor environments. This system enabled precise drone positioning indoors and utilized global and local path planners to generate flight paths that avoided dynamic, static, unknown, and known obstacles, demonstrating high practicality and feasibility. To achieve warehouse inventory inspection, a reinforcement learning approach was proposed, recognizing markers by adjusting the UAV’s viewpoint. We addressed several of the main problems in inventory management, including efficiently planning of paths, while ensuring a certain detection rate. Two reinforcement learning techniques, AC (actor–critic) and PPO (proximal policy optimization), were implemented based on AprilTag identification. Testing was performed in both simulated and real-world environments, and the effectiveness of the proposed method was validated. Full article
(This article belongs to the Section Drone Design and Development)
Show Figures

Figure 1

Figure 1
<p>The system architecture of our localization and navigation technique for UAVs. It consists of two components, indoor mapping and localization, and path planning with obstacle avoidance. In indoor mapping and localization, various global and local indoor positioning techniques, including UWB, QVIO, AprilTag, and RTAB-Map SLAM, are used for multi-sensor fusion.</p>
Full article ">Figure 2
<p>System flowchart of the proposed technique for indoor mapping and localization. After setting up the environment, the different positioning techniques were activated, and the pose obtained from each sensor was converted into a path. The paths were finally shown with the ROS visualization tool.</p>
Full article ">Figure 3
<p>The flowchart of our proposed navigation system. It also consists of loop closure detection, which helps to eliminate the errors accumulated in visual odometry over time.</p>
Full article ">Figure 4
<p>The architecture of the proposed warehouse inventory management system. The images acquired in the setup environment were used as inputs for training and testing. The data collection was divided into virtual and real scene cases. In the virtual environment, a simulated drone acquired images from various sampling angles of the shelves. In the real environment, images were taken from fixed points within the setup area using a camera mounted on a mobile platform. The testing process was identical to training, with the difference being that all parameters were fixed, and each step’s images were saved. Testing was conducted ten times, and the results are summarized in Table 4.</p>
Full article ">Figure 5
<p>A schematic diagram and internal structure of AC (actor–critic) framework. AC (actor-critic) is an on-policy method, and the basic framework is shown in (<b>a</b>). The internal structure of the actor is illustrated in (<b>b</b>).</p>
Full article ">Figure 6
<p>The framework for PPO (proximal policy optimization). It consists of two parts: an actor (<b>top</b>) and critic (<b>bottom</b>). There are two policies with the same structure but used for different purposes, one for interacting with the environment and the other for learning and updating.</p>
Full article ">Figure 7
<p>System flowchart of our reinforcement learning training framework. A set of random coordinates, angles, and images are taken as initial input. The system checks if sufficient data have been collected. If not, data collection continues. If enough data are collected, the learning process begins, and parameters are updated upon completion. Before feature extraction in the input training framework, images are normalized in size. The feature extraction network, ResNet34, is used, and its output is divided into two parts: policy and critic. The latter predicts an evaluation value, while the former outputs a probability distribution of actions. Action range checking is performed next, followed by selection of the next state and reward output according to the action selection rules. The system then checks if the task is complete, i.e., if all targets in the scene have been found. If not, a series of continuous state collections continues. Finally, the training results are observed based on the accumulated rewards from each round.</p>
Full article ">Figure 8
<p>The flowchart of our reinforcement learning architecture for testing. After training is completed, the trained parameters are saved and imported for testing. The testing was conducted using a test dataset as input and followed the same framework. Evaluation was based on the detection rate of objects and the number of steps used.</p>
Full article ">Figure 9
<p>The action movement in translation ((<b>a</b>), in front view) and rotation ((<b>b</b>), in top view). The choice of the next move included ‘up’, ‘down’, ‘left’, ‘right’, ‘upper-right’, ‘upper-left’, ’down-right’, and ‘down-left’. The camera orientation ranged from −40° to 40°, with an interval of 10°.</p>
Full article ">Figure 10
<p>The sampling points were derived from the viewpoints using polar coordinates <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>θ</mi> <mo>,</mo> <mi>ϕ</mi> <mo>)</mo> </mrow> </semantics></math>, where <span class="html-italic">p</span> is the sampling point, and <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>,</mo> <mi>θ</mi> <mo>,</mo> <mi>ϕ</mi> </mrow> </semantics></math> represent the radius, azimuth angle, and elevation angle, respectively.</p>
Full article ">Figure 11
<p>The navigation system architecture. It contained an UAV equipped with an Intel RealSense D435 depth camera and an UP-Board embedded computer.</p>
Full article ">Figure 12
<p>The positioning system architecture. A mobile platform was used to carry the UAV along with the UWB Tag and the D455 depth camera.</p>
Full article ">Figure 13
<p>RTAB-Map—Real environment mapping. The mapping process included the generation of a 3D point cloud for visual feature loop closure detection, a 2D occupancy grid for global path planning and obstacle avoidance, and a 2D map showing the safe flying zone for the UAV.</p>
Full article ">Figure 14
<p>Comparison of the paths created with or without continuous AprilTag detection.</p>
Full article ">Figure 15
<p>Indoor localization trajectories obtained from the RTAB-Map SLAM mapping and localization modes. The blue, green, and red paths represent the trajectories obtained when the initial mapping mode, localization mode, and the mapping mode were reactivated during the second localization mode.</p>
Full article ">Figure 16
<p>The characteristics of the different localization techniques. The yellow, cyan, red, and blue curves represent the ground truth, UWB, QVIO, and AprilTag localization trajectories, respectively.</p>
Full article ">Figure 17
<p>Path planning for dynamic obstacle avoidance—virtual environment. The red arrow represents the destination that the drone was heading in towards in RViz, the green path is the global path generated by the navfn global planner, and the yellow path is the local path generated by the local planner.</p>
Full article ">Figure 18
<p>Path planning for dynamic obstacle avoidance—real-world environment. The performance of the local path planner was evaluated using a puzzle made of foam blocks as static unknown obstacles.</p>
Full article ">Figure 19
<p>The boxes were placed on a shelf, and the drone equipped with a camera flew at a fixed height. There were three tracks in front of the shelf, each track had 35 sampling points, and the plane had a total of <math display="inline"><semantics> <mrow> <mn>35</mn> <mo>×</mo> <mn>3</mn> </mrow> </semantics></math> sampling angle points.</p>
Full article ">Figure 20
<p>The reinforcement learning training results for PPO and AC.</p>
Full article ">Figure 21
<p>The testing results using PPO in a virtual environment with step set to 50. The images on the right are zoomed-in to better illustrate the blue circles.</p>
Full article ">Figure 22
<p>The testing results using AC in a virtual scene with step set to 80.</p>
Full article ">Figure 23
<p>The real-world testing results for PPO training. The steps are from 1 to 4 and 10 to 13 in the left and right columns, respectively.</p>
Full article ">
15 pages, 2061 KiB  
Article
A Vision/Inertial Navigation/Global Navigation Satellite Integrated System for Relative and Absolute Localization in Land Vehicles
by Yao Zhang, Liang Chu, Yabin Mao, Xintong Yu, Jiawei Wang and Chong Guo
Sensors 2024, 24(10), 3079; https://doi.org/10.3390/s24103079 - 12 May 2024
Viewed by 984
Abstract
This paper presents an enhanced ground vehicle localization method designed to address the challenges associated with state estimation for autonomous vehicles operating in diverse environments. The focus is specifically on the precise localization of position and orientation in both local and global coordinate [...] Read more.
This paper presents an enhanced ground vehicle localization method designed to address the challenges associated with state estimation for autonomous vehicles operating in diverse environments. The focus is specifically on the precise localization of position and orientation in both local and global coordinate systems. The proposed approach integrates local estimates generated by existing visual–inertial odometry (VIO) methods into global position information obtained from the Global Navigation Satellite System (GNSS). This integration is achieved through optimizing fusion in a pose graph, ensuring precise local estimation and drift-free global position estimation. Considering the inherent complexities in autonomous driving scenarios, such as the potential failures of a visual–inertial navigation system (VINS) and restrictions on GNSS signals in urban canyons, leading to disruptions in localization outcomes, we introduce an adaptive fusion mechanism. This mechanism allows seamless switching between three modes: utilizing only VINS, using only GNSS, and normal fusion. The effectiveness of the proposed algorithm is demonstrated through rigorous testing in the Carla simulation environment and challenging UrbanNav scenarios. The evaluation includes both qualitative and quantitative analyses, revealing that the method exhibits robustness and accuracy. Full article
(This article belongs to the Section Navigation and Positioning)
Show Figures

Figure 1

Figure 1
<p>The scheme of proposed system.</p>
Full article ">Figure 2
<p>Diagram of the coordinates involved.</p>
Full article ">Figure 3
<p>Pose graph structure.</p>
Full article ">Figure 4
<p>The trajectories in Carla. <b>Truth_traj</b> represents the ground truth trajectory, <b>absolute</b> is the trajectory obtained from the proposed adaptive algorithm, and <b>relative</b> is the trajectory obtained from the proposed two-stage initialization process. <b>Vins_fusion</b> is the trajectory obtained from VINS-FUSION (without GNSS). <b>Vins_fusion_global</b> is the trajectory obtained from VINS-FUSION (with GNSS).</p>
Full article ">Figure 5
<p>The trajectories in the UrbanNav-HK-Data20190314. (<b>a</b>) shows the result trajectories obtained from our proposed algorithm compared to VINS-FUSION (with GNSS), <b>absolute</b> is the trajectory obtained from the proposed adaptive algorithm, <b>and vins_fusion_global</b> is the trajectory obtained from the VINS-FUSION; (<b>b</b>) shows the result trajectories obtained when our method and VINS-FUSION were extended to VINS-MONO. <b>adaptive_vins_mono</b> is the trajectory obtained from the proposed adaptive algorithm, and <b>vins_mono_global</b> is the trajectory obtained from VINS-FUSION. The red box indicates visual–inertial odometry (VIO) failure.</p>
Full article ">Figure 6
<p>Different system trajectories’ APE compared to ground truth. The x-axis, “Distances”, represents the distance from the starting point in meters.</p>
Full article ">
24 pages, 2438 KiB  
Review
Visual SLAM for Unmanned Aerial Vehicles: Localization and Perception
by Licong Zhuang, Xiaorong Zhong, Linjie Xu, Chunbao Tian and Wenshuai Yu
Sensors 2024, 24(10), 2980; https://doi.org/10.3390/s24102980 - 8 May 2024
Viewed by 1536
Abstract
Localization and perception play an important role as the basis of autonomous Unmanned Aerial Vehicle (UAV) applications, providing the internal state of movements and the external understanding of environments. Simultaneous Localization And Mapping (SLAM), one of the critical techniques for localization and perception, [...] Read more.
Localization and perception play an important role as the basis of autonomous Unmanned Aerial Vehicle (UAV) applications, providing the internal state of movements and the external understanding of environments. Simultaneous Localization And Mapping (SLAM), one of the critical techniques for localization and perception, is facing technical upgrading, due to the development of embedded hardware, multi-sensor technology, and artificial intelligence. This survey aims at the development of visual SLAM and the basis of UAV applications. The solutions to critical problems for visual SLAM are shown by reviewing state-of-the-art and newly presented algorithms, providing the research progression and direction in three essential aspects: real-time performance, texture-less environments, and dynamic environments. Visual–inertial fusion and learning-based enhancement are discussed for UAV localization and perception to illustrate their role in UAV applications. Subsequently, the trend of UAV localization and perception is shown. The algorithm components, camera configuration, and data processing methods are also introduced to give comprehensive preliminaries. In this paper, we provide coverage of visual SLAM and its related technologies over the past decade, with a specific focus on their applications in autonomous UAV applications. We summarize the current research, reveal potential problems, and outline future trends from academic and engineering perspectives. Full article
Show Figures

Figure 1

Figure 1
<p>General keyframe-based visual SLAM algorithm pipeline (similar to [<a href="#B22-sensors-24-02980" class="html-bibr">22</a>,<a href="#B23-sensors-24-02980" class="html-bibr">23</a>,<a href="#B24-sensors-24-02980" class="html-bibr">24</a>]). ‘Image Processing’ includes image distortion, feature extraction, etc., and procedures for image matching to find correspondence; ’Keyframe Decision’ considers whether the current map supports odometry to estimate the current state; ’Map generation’ triangulates the pixels to 3D map points; ’History updating’ updates the dataset by inserting new keyframes for potential loop detection.</p>
Full article ">Figure 2
<p>Finding correspondences between the current frame and the reference for the feature-based method (images come from EuRoC [<a href="#B41-sensors-24-02980" class="html-bibr">41</a>] dataset). Detection detects pixels with distinctiveness and repeatability; description creates the unique descriptor of features for feature matching; matching compares the similarity of those descriptors to match features.</p>
Full article ">Figure 3
<p>IMU measurements pre-integration: letters a and w, respectively, represent acceleration and angular velocity; dP, dQ, and dV are increments of position, orientation, and velocity between the last frame to the current frame.</p>
Full article ">Figure 4
<p>Different type of maps from EuRoC [<a href="#B41-sensors-24-02980" class="html-bibr">41</a>] dataset.</p>
Full article ">Figure 5
<p>Three schemes to build a dense map (experiment dataset [<a href="#B119-sensors-24-02980" class="html-bibr">119</a>]): (a) and (b) use different ways to produce depth image, the former measuring depth by depth sensors, while the latter estimates depth by deep learning modules and reconstructs a dense map by retrieving the pose of the depth images; (c) a NeRF-based SLAM produces a NeRF representation of environments. Other dense SLAM algorithms usually result in poor performance because of the lack of joint optimization.</p>
Full article ">
22 pages, 3441 KiB  
Article
Autonomous Navigation UAVs for Enhancing Information Freshness for Reconnaissance
by Chen Xie, Binbin Wu, Daoxing Guo and Xiao Chen
Electronics 2024, 13(7), 1354; https://doi.org/10.3390/electronics13071354 - 3 Apr 2024
Viewed by 763
Abstract
The demand for autonomous navigation UAVs in reconnaissance is steadily increasing. One crucial aspect of these missions is the freshness of reconnaissance information, which serves as a vital indicator of mission effectiveness. However, there is a lack of targeted investigation in the research [...] Read more.
The demand for autonomous navigation UAVs in reconnaissance is steadily increasing. One crucial aspect of these missions is the freshness of reconnaissance information, which serves as a vital indicator of mission effectiveness. However, there is a lack of targeted investigation in the research on autonomous single/multi-UAV missions and joint path planning. Furthermore, the use of visual–inertial odometry (VIO) in rotary-wing UAVs can lead to significant positional drift during flight, which may result in the loss of the UAV and mission failure. This paper investigates joint planning problems in single/multi-UAV reconnaissance missions under both GPS-available and GPS-unavailable scenarios and proposes an integrated data collection and beacon-assisted localization approach. Finally, the numerical results demonstrate the effectiveness of the proposed scheme in enhancing the freshness of reconnaissance information. Full article
Show Figures

Figure 1

Figure 1
<p>The AoI growth rates of different types of reconnaissance nodes.</p>
Full article ">Figure 2
<p>AoI model for reconnaissance nodes with heterogeneous urgency.</p>
Full article ">Figure 3
<p>The integrated solution for reconnaissance data collection and beacon-assisted positioning, which includes visual beacon positioning correction (<b>left</b>) and wireless beacon positioning correction (<b>right</b>).</p>
Full article ">Figure 4
<p>Nodes and mission environment schematic diagram, along with corresponding three-dimensional topological roadmap. (<b>a</b>) Random distribution of 15 nodes. (<b>b</b>) Topological roadmap of 15 nodes. (<b>c</b>) Random distribution of 30 nodes. (<b>d</b>) Topological roadmap of 30 nodes. (<b>e</b>) Random distribution of 45 nodes. (<b>f</b>) Topological roadmap of 45 nodes. (<b>g</b>) Circular distribution of 20 nodes. (<b>h</b>) Topological roadmap of 20 nodes.</p>
Full article ">Figure 5
<p>The comparative results of 30 Monte Carlo simulations between the direct optimization scheme and the minimum total flight time scheme in the scenario of ground reconnaissance nodes with a regular distribution, where the red crosses represent outliers caused by the optimization algorithm getting trapped in local optima. (<b>a</b>) A box plot comparing the AAoI metric. (<b>b</b>) A box plot comparing the total flight time metric.</p>
Full article ">Figure 6
<p>A comparison of results with target values near the median. (<b>a</b>) The path planned for the direct optimization scheme. (<b>b</b>) The path planned for the total flight time minimization scheme.</p>
Full article ">Figure 7
<p>Comparing the results of 30 Monte Carlo simulations between the direct optimization scheme and the total flight time minimization scheme in a scenario where ground reconnaissance nodes are randomly distributed. (<b>a</b>) A box plot comparing the AAoI metric in a scenario with 15 nodes. (<b>b</b>) A box plot comparing the total flight time metric in a scenario with 15 nodes. (<b>c</b>) A box plot comparing the AAoI metric in a scenario with 30 nodes. (<b>d</b>) A box plot comparing the total flight time metric in a scenario with 30 nodes. (<b>e</b>) A box plot comparing the AAoI metric in a scenario with 45 nodes. (<b>f</b>) A box plot comparing the total flight time metric in a scenario with 45 nodes.</p>
Full article ">Figure 8
<p>A comparison of results with target values near the median. (<b>a</b>) The planned path of the direct optimization scheme in the scenario with 15 nodes. (<b>b</b>) The planned path of the total flight time minimization scheme in the scenario with 15 nodes. (<b>c</b>) The planned path of the direct optimization scheme in the scenario with 30 nodes. (<b>d</b>) The planned path of the total flight time minimization scheme in the scenario with 30 nodes. (<b>e</b>) The planned path of the direct optimization scheme in the scenario with 45 nodes. (<b>f</b>) The planned path of the total flight time minimization scheme in the scenario with 45 nodes.</p>
Full article ">Figure 9
<p>Comparing the results of 30 Monte Carlo simulations between the combined scheme and the total flight time minimization scheme in the VIO + data collection and beacon-assisted integrated navigation mode. (<b>a</b>) A box plot comparing the AAoI metric in a scenario with 30 nodes. (<b>b</b>) A box plot comparing the total flight time metric in a scenario with 30 nodes. (<b>c</b>) A box plot comparing the AAoI metric in a scenario with 45 nodes. (<b>d</b>) A box plot comparing the total flight time metric in a scenario with 45 nodes.</p>
Full article ">Figure 10
<p>Comparing the results of 30 Monte Carlo simulations for the direct optimization scheme, minimizing total flight time scheme, and minimizing total mission duration scheme in the VIO + GPS navigation mode. (<b>a</b>) A box plot comparing the AAoI metric in multi-UAV reconnaissance. (<b>b</b>) A box plot comparing the total flight time metric in multi-UAV reconnaissance. (<b>c</b>) A box plot comparing the maximum flight time metric in multi-UAV reconnaissance.</p>
Full article ">Figure 11
<p>A comparison of results with target values near the median. The paths of UAVs corresponding to five different types of ground nodes are represented by lines of respective colors. Meanwhile, the paths of UAVs corresponding to all types of ground nodes are represented by light blue lines. (<b>a</b>) The planned path of the direct optimization scheme. (<b>b</b>) The planned path of the minimizing total flight time scheme. (<b>c</b>) The planned path of the minimizing maximum flight time scheme.</p>
Full article ">Figure 12
<p>The Monte Carlo simulation results for the integrated VIO + data collection and beacon-assisted localization navigation mode, comparing the minimizing maximum flight time scheme, joint scheme 1, and joint scheme 4.</p>
Full article ">
25 pages, 14610 KiB  
Article
Improving SLAM Techniques with Integrated Multi-Sensor Fusion for 3D Reconstruction
by Yiyi Cai, Yang Ou and Tuanfa Qin
Sensors 2024, 24(7), 2033; https://doi.org/10.3390/s24072033 - 22 Mar 2024
Cited by 3 | Viewed by 2015
Abstract
Simultaneous Localization and Mapping (SLAM) poses distinct challenges, especially in settings with variable elements, which demand the integration of multiple sensors to ensure robustness. This study addresses these issues by integrating advanced technologies like LiDAR-inertial odometry (LIO), visual-inertial odometry (VIO), and sophisticated Inertial [...] Read more.
Simultaneous Localization and Mapping (SLAM) poses distinct challenges, especially in settings with variable elements, which demand the integration of multiple sensors to ensure robustness. This study addresses these issues by integrating advanced technologies like LiDAR-inertial odometry (LIO), visual-inertial odometry (VIO), and sophisticated Inertial Measurement Unit (IMU) preintegration methods. These integrations enhance the robustness and reliability of the SLAM process for precise mapping of complex environments. Additionally, incorporating an object-detection network aids in identifying and excluding transient objects such as pedestrians and vehicles, essential for maintaining the integrity and accuracy of environmental mapping. The object-detection network features a lightweight design and swift performance, enabling real-time analysis without significant resource utilization. Our approach focuses on harmoniously blending these techniques to yield superior mapping outcomes in complex scenarios. The effectiveness of our proposed methods is substantiated through experimental evaluation, demonstrating their capability to produce more reliable and precise maps in environments with variable elements. The results indicate improvements in autonomous navigation and mapping, providing a practical solution for SLAM in challenging and dynamic settings. Full article
(This article belongs to the Section Sensors and Robotics)
Show Figures

Figure 1

Figure 1
<p>YOLOv5 algorithm network structure.</p>
Full article ">Figure 2
<p>Multi-sensor tightly coupled system architecture.</p>
Full article ">Figure 3
<p>Schematic representation of the sub-system: visual–inertial integration.</p>
Full article ">Figure 4
<p>Schematic representation of the sub-system: LiDAR–inertial integration.</p>
Full article ">Figure 5
<p>Depth correlation.</p>
Full article ">Figure 6
<p>The reconstructed map of the “hku_main_building” sequence: (<b>a</b>) is a bird’s-eye view, capturing the overall structure. (<b>b</b>,<b>c</b>) are closeups, revealing intricate details.</p>
Full article ">Figure 7
<p>The reconstructed map of the “hku_campus_seq_03” sequence: (<b>a</b>) offers a bird’s-eye view of the entire radiance map, providing a comprehensive overview of the campus. (<b>b</b>,<b>c</b>) then zoom in to showcase the intricate details of specific areas.</p>
Full article ">Figure 8
<p>The reconstructed map of the “hku_park_00” sequence: (<b>a</b>) Presents a bird’s-eye view, capturing the park’s expanse. Detailed closeups are shown in (<b>b</b>,<b>c</b>), offering a deeper glimpse into its features.</p>
Full article ">Figure 9
<p>The reconstructed map of the “hkust_campus_02” sequence: (<b>a</b>) Presents a bird’s-eye view of a specific area of campus, providing an overview of its layout. For a deeper understanding of its intricate details, refer to the closeup views in (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 10
<p>The reconstructed map of the “degenerate_seq_02” sequence: (<b>a</b>) Offers a bird’s-eye view of the corridor, outlining its structure. For a closer look at the intricate details within, refer to the closeup views presented in (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 11
<p>(<b>a</b>–<b>f</b>) Trajectory comparison: our system vs. ground truth.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>d</b>) Trajectory comparison: our system vs. LIO-SAM.</p>
Full article ">
20 pages, 3494 KiB  
Article
Visual–Inertial Odometry of Structured and Unstructured Lines Based on Vanishing Points in Indoor Environments
by Xiaojing He, Baoquan Li, Shulei Qiu and Kexin Liu
Appl. Sci. 2024, 14(5), 1990; https://doi.org/10.3390/app14051990 - 28 Feb 2024
Viewed by 817
Abstract
In conventional point-line visual–inertial odometry systems in indoor environments, consideration of spatial position recovery and line feature classification can improve localization accuracy. In this paper, a monocular visual–inertial odometry based on structured and unstructured line features of vanishing points is proposed. First, the [...] Read more.
In conventional point-line visual–inertial odometry systems in indoor environments, consideration of spatial position recovery and line feature classification can improve localization accuracy. In this paper, a monocular visual–inertial odometry based on structured and unstructured line features of vanishing points is proposed. First, the degeneracy phenomenon caused by a special geometric relationship between epipoles and line features is analyzed in the process of triangulation, and a degeneracy detection strategy is designed to determine the location of the epipoles. Then, considering that the vanishing point and the epipole coincide at infinity, the vanishing point feature is introduced to solve the degeneracy and direction vector optimization problem of line features. Finally, threshold constraints are used to categorize straight lines into structural and non-structural features under the Manhattan world assumption, and the vanishing point measurement model is added to the sliding window for joint optimization. Comparative tests on the EuRoC and TUM-VI public datasets validated the effectiveness of the proposed method. Full article
(This article belongs to the Topic Multi-Sensor Integrated Navigation Systems)
Show Figures

Figure 1

Figure 1
<p>Block diagram for the proposed method.</p>
Full article ">Figure 2
<p>Illustrative diagram of degeneracy phenomenon of line features.</p>
Full article ">Figure 3
<p>Analytical diagram of the relationship between spatial positions of the vanishing point and the epipole.</p>
Full article ">Figure 4
<p>Schematic diagram of the residuals at vanishing point.</p>
Full article ">Figure 5
<p>The experiment snapshots of the EuRoC dataset by the proposed method.</p>
Full article ">Figure 6
<p>Evolution of localization with respect to the V1_02_medium sequence (blue solid lines: the proposed method; blue dashed lines: PL−VINS; green solid lines: UV−SLAM; red solid lines: VINS−Mono; and red thick dashed lines: ground truth).</p>
Full article ">Figure 7
<p>Comparative paths with respect to V1_02_medium (<b>left</b>) and MH_02_easy (<b>right</b>) (blue solid lines: the proposed method; blue dashed lines: PL−VINS; green solid lines: UV−SLAM; red solid lines: VINS−Mono; and red thick dashed lines: ground truth).</p>
Full article ">Figure 8
<p>Comparison of the pose trajectory error of UV-SLAM (<b>left</b>) and the proposed method (<b>right</b>) (top plots: MH_03_medium; middle plots: V1_03_ difficult; bottom plots: V2_01_easy; left: UV-SLAM; and right: the proposed method).</p>
Full article ">Figure 9
<p>Comparative paths with respect to room2_512_16 (<b>left</b>) and room5_512_16 (<b>right</b>) (blue dashed lines: the proposed method; green solid lines: UV−SLAM; red solid lines: VINS−Mono; and red thick dashed lines: ground truth).</p>
Full article ">Figure 10
<p>APE evolution results regarding VINS-Mono (<b>left</b>) and the proposed method (<b>right</b>) experiments (<b>top</b> plots: room1_512_16; <b>middle</b> plots: room3_512_16; and <b>bottom</b> plots: room5_512_16).</p>
Full article ">
19 pages, 10597 KiB  
Article
Enhanced Seamless Indoor–Outdoor Tracking Using Time Series of GNSS Positioning Errors
by Eduard Angelats, Alban Gorreja, Pedro F. Espín-López, M. Eulàlia Parés, Eva Savina Malinverni and Roberto Pierdicca
ISPRS Int. J. Geo-Inf. 2024, 13(3), 72; https://doi.org/10.3390/ijgi13030072 - 27 Feb 2024
Viewed by 1743
Abstract
The seamless integration of indoor and outdoor positioning has gained considerable attention due to its practical implications in various fields. This paper presents an innovative approach aimed at detecting and delineating outdoor, indoor, and transition areas using a time series analysis of Global [...] Read more.
The seamless integration of indoor and outdoor positioning has gained considerable attention due to its practical implications in various fields. This paper presents an innovative approach aimed at detecting and delineating outdoor, indoor, and transition areas using a time series analysis of Global Navigation Satellite System (GNSS) error statistics. By leveraging this contextual understanding, the decision-making process between GNSS-based and Visual-Inertial Odometry (VIO) for trajectory estimation is refined, enabling a more robust and accurate positioning. The methodology involves three key steps: proposing the division of our context environment into a set of areas (indoor, outdoor, and transition), exploring two methodologies for the classification of space based on a time series of GNSS error statistics, and refining the trajectory estimation strategy based on contextual knowledge. Real data across diverse scenarios validate the approach, yielding trajectory estimations with accuracy consistently below 10 m. Full article
(This article belongs to the Topic Urban Sensing Technologies)
Show Figures

Figure 1

Figure 1
<p>Positioning system mounted on a helmet designed and built by CTTC.</p>
Full article ">Figure 2
<p>Space definition proposed by [<a href="#B31-ijgi-13-00072" class="html-bibr">31</a>] and adapted by [<a href="#B22-ijgi-13-00072" class="html-bibr">22</a>].</p>
Full article ">Figure 3
<p>Left side: sky representation with projected satellites (green = LOS; red = NLOS). Right side: schemes of the multipath and NLOS effect [<a href="#B32-ijgi-13-00072" class="html-bibr">32</a>].</p>
Full article ">Figure 4
<p>Flowchart depicting the algorithm used to determine indoor, outdoor, and transition areas using the GNSS error estimation.</p>
Full article ">Figure 5
<p>Flowchart depicting the algorithm used to determine indoor, outdoor, and transition areas using the slope approach.</p>
Full article ">Figure 6
<p>Flowchart depicting the algorithm used to determine indoor, outdoor, and transition areas using the MK approach.</p>
Full article ">Figure 7
<p>Example of an area next to a building marked as transition or semi-outdoor area.</p>
Full article ">Figure 8
<p>Transition or semi-outdoor area for the “PMT-LAB” (red) and GNSS-based trajectory (green). Building layouts are shown in blue.</p>
Full article ">Figure 9
<p>GNSS trajectory horizontal error vs. time. Vertical lines highlight relevant events such as entering or going out from transitional areas and buildings. Horizontal orange line shows a a line with constant horizontal error of 5 m, while the horizontal green line shows of 3.5 m.</p>
Full article ">Figure 10
<p>GNSS trajectory vertical error vs. time.</p>
Full article ">Figure 11
<p>Slope calculated in a time window of 10 s, using as input the horizontal error (red) and vertical error (blue).</p>
Full article ">Figure 12
<p>Slope calculated in a time window of 5, using as input the horizontal error (red) and vertical error (blue).</p>
Full article ">Figure 13
<p>Slope calculated in a time window of 3 s, using as input the horizontal error (red) and vertical error (blue).</p>
Full article ">Figure 14
<p>Slope test applied to the trajectory of “PMT-LAB”. Yellow points are points classified as good GNSS areas, while green points are points belonging to be transition or indoor areas. Building layouts are shown in blue, while transition or semi-outdoor area for the “PMT-LAB” is shown in red.</p>
Full article ">Figure 15
<p>MK test applied to the trajectory of “PMT-LAB”. Green points are good GNSS areas, orange points are transitional areas from outdoor to indoor, and red are from indoor to outdoor. Building layouts are shown in blue, while transition or semi-outdoor area for the “PMT-LAB” is shown in red.</p>
Full article ">Figure 16
<p>MK test applied to the trajectory in Garraf town. Green points are good GNSS areas, orange points are transitional areas from outdoor to indoor. Building layouts are shown in blue while transition or semi-outdoor area for the Garraf town is shown in red.</p>
Full article ">Figure 17
<p>MK test applied to the trajectory of Falconera (<b>left</b>) and Colluspina (<b>right</b>). On the left figure, green points are good GNSS areas, orange are transitional areas from outdoor to indoor, and red are from indoor to outdoor. On the right, blue points are good GNSS areas while red points are indoor areas.</p>
Full article ">Figure 18
<p>Comparison of the older seamless indoor–outdoor trajectory (red) with the newer one (blue) for the “PMT-LAB”. Ground truth trajectory is depicted in black (dashed line).</p>
Full article ">Figure 19
<p>Comparison of the older (red) estimated trajectory with the newer one (blue) in Collsuspina site. Ground truth trajectory is depicted in black (dashed line).</p>
Full article ">Figure 20
<p>The final segments of the Falconera and Garraf estimated trajectories conclude at the same building. The left side displays the previous trajectories, while the right side shows the updated trajectories.</p>
Full article ">
15 pages, 1559 KiB  
Article
LRPL-VIO: A Lightweight and Robust Visual–Inertial Odometry with Point and Line Features
by Feixiang Zheng, Lu Zhou, Wanbiao Lin, Jingyang Liu and Lei Sun
Sensors 2024, 24(4), 1322; https://doi.org/10.3390/s24041322 - 18 Feb 2024
Cited by 2 | Viewed by 1682
Abstract
Visual-inertial odometry (VIO) algorithms, fusing various features such as points and lines, are able to improve their performance in challenging scenes while the running time severely increases. In this paper, we propose a novel lightweight point–line visual–inertial odometry algorithm to solve this problem, [...] Read more.
Visual-inertial odometry (VIO) algorithms, fusing various features such as points and lines, are able to improve their performance in challenging scenes while the running time severely increases. In this paper, we propose a novel lightweight point–line visual–inertial odometry algorithm to solve this problem, called LRPL-VIO. Firstly, a fast line matching method is proposed based on the assumption that the photometric values of endpoints and midpoints are invariant between consecutive frames, which greatly reduces the time consumption of the front end. Then, an efficient filter-based state estimation framework is designed to finish information fusion (point, line, and inertial). Fresh measurements of line features with good tracking quality are selected for state estimation using a unique feature selection scheme, which improves the efficiency of the proposed algorithm. Finally, validation experiments are conducted on public datasets and in real-world tests to evaluate the performance of LRPL-VIO and the results show that we outperform other state-of-the-art algorithms especially in terms of speed and robustness. Full article
(This article belongs to the Section Sensors and Robotics)
Show Figures

Figure 1

Figure 1
<p>The working flowchart of LRPL-VIO.</p>
Full article ">Figure 2
<p>The pose estimation error of PL-VINS and LRPL-VIO on the UMA-VI and VIODE dataset. (<b>a</b>) The alignment error of PL-VINS in class_csc2. (<b>b</b>) The alignment error of PL-VINS in parking_csc2. (<b>c</b>) The RMSE ATE of PL-VINS in cd3_high. (<b>d</b>) The alignment error of LRPL-VIO in class_csc2. (<b>e</b>) The alignment error of LRPL-VIO in parking_csc2. (<b>f</b>) The RMSE ATE of LRPL-VIO in cd3_high.</p>
Full article ">Figure 3
<p>The figures of real-world experiments. (<b>a</b>) An example image of sequence Lab_Normal. (<b>b</b>) An example image of sequence Lab_FastRotation. (<b>c</b>) The 3D error map of HybVIO in Lab_Normal. (<b>d</b>) The X-Y plane of 3D error map of HybVIO in Lab_Normal. (<b>e</b>) The 3D error map of HybVIO in Lab_FastRotation. (<b>f</b>) The X-Y plane of 3D error map of HybVIO in Lab_FastRotation. (<b>g</b>) The 3D error map of LRPL-VIO in Lab_Normal. (<b>h</b>) The X-Y plane of 3D error map of LRPL-VIO in Lab_Normal. (<b>i</b>) The 3D error map of LRPL-VIO in Lab_FastRotation. (<b>j</b>) The X-Y plane of 3D error map of LRPL-VIO in Lab_FastRotation.</p>
Full article ">
Back to TopTop