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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (552)

Search Parameters:
Keywords = inverse kinematics

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 521 KiB  
Review
A Review on Inverse Kinematics, Control and Planning for Robotic Manipulators With and Without Obstacles via Deep Neural Networks
by Ana Calzada-Garcia, Juan G. Victores, Francisco J. Naranjo-Campos and Carlos Balaguer
Algorithms 2025, 18(1), 23; https://doi.org/10.3390/a18010023 (registering DOI) - 4 Jan 2025
Viewed by 457
Abstract
Robotic manipulators are highly valuable tools that have become widespread in the industry, as they can achieve great precision and velocity in pick and place as well as processing tasks. However, to unlock their complete potential, some problems such as inverse kinematics (IK) [...] Read more.
Robotic manipulators are highly valuable tools that have become widespread in the industry, as they can achieve great precision and velocity in pick and place as well as processing tasks. However, to unlock their complete potential, some problems such as inverse kinematics (IK) need to be solved: given a Cartesian target, a method is needed to find the right configuration for the robot to reach that point. Another issue that needs to be addressed when dealing with robotic manipulators is the obstacle avoidance problem. Workspaces are usually cluttered and the manipulator should be able to avoid colliding with objects that could damage it, as well as with itself. Two alternatives exist to do this: a controller can be designed that computes the best action for each moment given the manipulator’s state, or a sequence of movements can be planned to be executed by the robot. Classical approaches to all these problems, such as numeric or analytical methods, can produce precise results but take a high computation time and do not always converge. Learning-based methods have gained considerable attention in tackling the IK problem, as well as motion planning and control. These methods can reduce the computational cost and provide results for every situation avoiding singularities. This article presents a literature review of the advances made in the past five years in the use of Deep Neural Networks (DNN) for IK with regard to control and planning with and without obstacles for rigid robotic manipulators. The literature has been organized in several categories depending on the type of DNN used to solve the problem. The main contributions of each reference are reviewed and the best results are presented in summary tables. Full article
(This article belongs to the Special Issue Optimization Methods for Advanced Manufacturing)
Show Figures

Figure 1

Figure 1
<p>Fully connected feed-forward Deep Neural Network for learning inverse kinematics.</p>
Full article ">Figure 2
<p>Fully connected feed-forward Deep Neural Network for learning control with obstacles.</p>
Full article ">Figure 3
<p>Fully connected feed-forward Deep Neural Network for learning motion planning with obstacles.</p>
Full article ">
19 pages, 313 KiB  
Article
A Closed Form of Higher-Order Cayley Transforms and Generalized Rodrigues Vectors Parameterization of Rigid Motion
by Daniel Condurache, Mihail Cojocari and Ioan-Adrian Ciureanu
Mathematics 2025, 13(1), 114; https://doi.org/10.3390/math13010114 - 30 Dec 2024
Viewed by 276
Abstract
This paper introduces a novel closed-form coordinate-free expression for the higher-order Cayley transform, a concept that has not been explored in depth before. The transform is defined by the Lie algebra of three-dimensional vectors into the Lie group of proper orthogonal Euclidean tensors. [...] Read more.
This paper introduces a novel closed-form coordinate-free expression for the higher-order Cayley transform, a concept that has not been explored in depth before. The transform is defined by the Lie algebra of three-dimensional vectors into the Lie group of proper orthogonal Euclidean tensors. The approach uses only elementary algebraic calculations with Euclidean vectors and tensors. The analytical expressions are given by rational functions by the Euclidean norm of vector parameterization. The inverse of the higher-order Cayley map is a multi-valued function that recovers the higher-order Rodrigues vectors (the principal parameterization and their shadows). Using vector parameterizations of the Euler and higher-order Rodrigues vectors, we determine the instantaneous angular velocity (in space and body frame), kinematics equations, and tangent operator. The analytical expressions of the parameterized quantities are identical for both the principal vector and shadows parameterization, showcasing the novelty and potential of our research. Full article
(This article belongs to the Special Issue Geometric Methods in Contemporary Engineering)
18 pages, 11046 KiB  
Article
Inverse and Forward Kinematics and CAD-Based Simulation of a 5-DOF Delta-Type Parallel Robot with Actuation Redundancy
by Pavel Laryushkin, Anton Antonov, Alexey Fomin and Oxana Fomina
Robotics 2025, 14(1), 1; https://doi.org/10.3390/robotics14010001 - 27 Dec 2024
Viewed by 393
Abstract
This article introduces a novel modification of a Delta-type parallel robot. The robot has five degrees of freedom and provides its end-effector with a 3T2R motion pattern (three translational and two rotational degrees of freedom). The fifth degree of freedom (rotation) is kinematically [...] Read more.
This article introduces a novel modification of a Delta-type parallel robot. The robot has five degrees of freedom and provides its end-effector with a 3T2R motion pattern (three translational and two rotational degrees of freedom). The fifth degree of freedom (rotation) is kinematically decoupled from the other four motions, and it is controlled by two drives. Thus, the proposed robot has a redundant actuation. In this article, we present an algorithm to solve the inverse kinematics of this robot and apply it to a path modeling example of a spiral-like trajectory. Numerical simulations illustrate the algorithm and show how the actuated coordinates change along the considered trajectory. Forward kinematics follows next, and an approach is introduced to determine all end-effector configurations for the specified displacements in the actuated joints. A numerical example presents four assembly modes of the robot corresponding to four real solutions of the forward kinematic problem. Finally, this article demonstrates a computer-aided design and analysis of the proposed robot: we describe a procedure for analyzing inverse kinematics and calculating actuation torques. This study forms the basis for the future manufacturing and experimental analysis of a robot prototype. Full article
(This article belongs to the Section Industrial Robots and Automation)
Show Figures

Figure 1

Figure 1
<p>Schematic design of the considered robot.</p>
Full article ">Figure 2
<p>Rotation angles in the auxiliary legs and tilt angles of the moving plate and end-effector.</p>
Full article ">Figure 3
<p>Two different initial configurations of the universal joint: (<b>a</b>) the yoke input shaft lies in the plane spanned by the input and output shafts; (<b>b</b>) the yoke input shaft is orthogonal to this plane.</p>
Full article ">Figure 4
<p>An example of two universal joints used consecutively.</p>
Full article ">Figure 5
<p>Simulated trajectory.</p>
Full article ">Figure 6
<p>The change of the end-effector coordinates along the simulated path.</p>
Full article ">Figure 7
<p>The change of the actuated coordinates along the simulated path.</p>
Full article ">Figure 8
<p>Four assembly modes of the robot that correspond to four solutions of the forward kinematics (<a href="#robotics-14-00001-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 9
<p>Computer model of the proposed 5-DOF Delta-type parallel robot with actuation redundancy: (<b>a</b>) isometric view with end-effector trajectory; (<b>b</b>) front view; (<b>c</b>) top view.</p>
Full article ">Figure 10
<p>Computed angular speeds (solid lines) and accelerations (dashed lines) of the motors.</p>
Full article ">Figure 11
<p>Robot links and their reference frames and axes for computing inertia parameters: (<b>a</b>) side carriage; (<b>b</b>) back carriage; (<b>c</b>) moving plate; (<b>d</b>) U-shaped link; (<b>e</b>) yoke; (<b>f</b>) end-effector; (<b>g</b>) screw; (<b>h</b>) U-rod; (<b>i</b>) S-rod.</p>
Full article ">Figure 12
<p>Computed motor torques.</p>
Full article ">
19 pages, 32077 KiB  
Article
Present-Day Tectonic Deformation Characteristics of the Northeastern Pamir Margin Constrained by InSAR and GPS Observations
by Junjie Zhang, Xiaogang Song, Donglin Wu and Xinjian Shan
Remote Sens. 2024, 16(24), 4771; https://doi.org/10.3390/rs16244771 - 21 Dec 2024
Viewed by 375
Abstract
The Pamir is located on the northwestern margin of the Tibetan Plateau, which is an area of intense continental deformation and part of the famous India–Himalaya collision zone. The dominant structural deformation in the eastern Pamir is characterized by a 250 km long [...] Read more.
The Pamir is located on the northwestern margin of the Tibetan Plateau, which is an area of intense continental deformation and part of the famous India–Himalaya collision zone. The dominant structural deformation in the eastern Pamir is characterized by a 250 km long east–west extensional fault system, known as the Kongur Shan extensional system (KSES), which has developed a series of faults with different orientations and characteristics, resulting in highly complex structural deformation and lacking sufficient geodetic constraints. We collected Sentinel-1 SAR data from December 2016 to March 2023, obtained high-resolution ascending and descending LOS velocities and 3D deformation fields, and combined them with GPS data to constrain the current motion characteristics of the northeastern Pamirs for the first time. Based on the two-dimensional screw dislocation model and using the Bayesian Markov chain Monte Carlo (MCMC) inversion method, the kinematic parameters of the fault were calculated, revealing the fault kinematic characteristics in this region. Our results demonstrate that the present-day deformation of the KSES is dominated by nearly E–W extension, with maximum extensional motion concentrated in its central segment, reaching peak extension rates of ~7.59 mm/yr corresponding to the Kongur Shan. The right-lateral Muji fault at the northern end exhibits equivalent rates of extensional motion with a relatively shallow locking depth. The strike-slip rate along the Muji fault gradually increases from west to east, ranging approximately between 4 and 6 mm/yr, significantly influenced by the eastern normal fault. The Tahman fault (TKF) at the southernmost end of the KSES shows an extension rate of ~1.5 mm/yr accompanied by minor strike-slip motion. The Kashi anticline is approaching stability, while the Mushi anticline along the eastern Pamir frontal thrust (PFT) remains active with continuous uplift at ~2 mm/yr, indicating that deformation along the Tarim Basin–Tian Shan boundary has propagated southward from the South Tian Shan thrust (STST). Overall, this study demonstrates the effectiveness of integrated InSAR and GPS data in constraining contemporary deformation patterns along the northeastern Pamir margin, contributing to our understanding of the region’s tectonic characteristics. Full article
Show Figures

Figure 1

Figure 1
<p>Tectonics and seismicity of the study area. (<b>a</b>) The yellow and red rectangle shows the spatial footprint of the Sentinel-1 InSAR coverage. Blue arrows show the GPS horizontal velocity field with respect to the stable Eurasian plate [<a href="#B28-remotesensing-16-04771" class="html-bibr">28</a>]. Circles of different colors represent earthquake events of varying magnitudes. (<b>b</b>) Fault structures in the eastern part of PFT. Red and pink focal mechanisms represent the mainshock and aftershocks of the 1985 Wuqia earthquake. Brown focal mechanisms represent the mainshock of the 2016 Aketao earthquake. (<b>c</b>) Fault segmentation in KESE. S1–S5 correspond to different segments, respectively. STST = southern Tian Shan thrust, PFT = Pamir frontal thrust, KSES = Kongur Shan extensional system, MPT = main Pamir thrust, KKF = Karakax fault, KYTS = Kashgar–Yecheng transfer system, TFF = Talas–Fergana fault, TT = Tuomuluoan thrust, MF = Muji fault, KATF = King Ata Tagh normal fault, KSF = Kongur Shan normal fault, MAF = Muztagh Ata normal fault, TF = Tahman normal fault, TKF = Tashkorgan normal fault.</p>
Full article ">Figure 2
<p>The Sentinel-1 A/B data processing workflow. It consists of three steps, including interferograms generation, SBAS time series analysis, and three-dimensional deformation field solution.</p>
Full article ">Figure 3
<p>Perpendicular and temporal baseline plot showing the network of interferograms on one ascending track (<b>a</b>) and one descending track (<b>b</b>) used in this study. The number of total interferograms are labelled for each track.</p>
Full article ">Figure 4
<p>Interpolated GNSS velocities map. The interpolated GNSS velocities using the method outlined by Shen et al. [<a href="#B47-remotesensing-16-04771" class="html-bibr">47</a>], (<b>a</b>) corresponds to EW and (<b>b</b>) corresponds to NS. GNSS velocities are resampled to a resolution of 0.01 degrees. Different colored circles represent different GPS data, and the color bars for GPS various data and the interpolated velocity field are identical.</p>
Full article ">Figure 5
<p>The satellite line-of-sight (LOS) velocity fields of the northeastern Pamir margin. Red lines represent the fault crossing profiles, each profile for 60 km long and 10 km wide, distributed along six sub-faults of the KSES. (<b>a</b>) corresponds to ascending track and (<b>b</b>) corresponds to descending track.</p>
Full article ">Figure 6
<p>Joint InSAR-GPS three-dimensional deformation field. (<b>a</b>–<b>c</b>) are east–west, north–south, and vertical components, respectively.</p>
Full article ">Figure 7
<p>GPS profiles and results. (<b>a</b>) Four profiles in KSES. Each profile is 300 km long and 50 km wide. (<b>b</b>) The GPS data was projected parallel to and perpendicular to the local fault, respectively, and combined with fault strike.</p>
Full article ">Figure 8
<p>The cross-fault profiles of ascending and descending LOS deformation velocities. (<b>a</b>) represents ascending LOS velocity profiles, and (<b>b</b>) represents descending LOS velocity profiles. (<b>a</b>–<b>f</b>) correspond successively to the six profiles in <a href="#remotesensing-16-04771-f005" class="html-fig">Figure 5</a>. Black dots are binned average values every 1 km along the profile. Gray vertical stripes indicate the mountains on profiles. Red and purple lines are the best-fitting models. The black dotted line indicates the fault location.</p>
Full article ">Figure 9
<p>An example of Bayesian MCMC inversion results for profile aa’. Posterior marginal probability density functions illustrating parameter estimation and uncertainty quantification. (<b>Top</b>): profile aa’ topography from the Copernicus DEM data with 30 m spatial resolution (average elevation: white line; min/max: gray lines). (<b>Middle</b>): InSAR LOS velocities with the best-fitting predicted velocities. (<b>Bottom</b>): model-predicted fault-parallel, fault-normal, and vertical velocities.</p>
Full article ">Figure 10
<p>The LOS velocities and vertical component of the eastern PFT. (<b>a</b>–<b>c</b>) are ascending, descending, and vertical components, respectively. The abnormal deformation area corresponding to the black circle and black rectangle are caused by industrial activity. The satellite images corresponding to these two regions are shown in <a href="#app1-remotesensing-16-04771" class="html-app">Figure S5</a>. (<b>d</b>–<b>f</b>) correspond to the results of profiles aa’, bb’, and cc’ in (<b>a</b>–<b>c</b>), respectively. The pink, blue, and green points correspond to the results of profiles aa’, bb’, and cc’, respectively.</p>
Full article ">
19 pages, 8083 KiB  
Article
Changes of Ankle Motion and Ground Reaction Force Using Elastic Neutral AFO in Neurological Patients with Inverted Foot During Gait
by Du-Jin Park and Young-In Hwang
Actuators 2024, 13(12), 526; https://doi.org/10.3390/act13120526 - 20 Dec 2024
Viewed by 343
Abstract
Many stroke patients develop ankle deformities due to neurological or non-neurological factors, resulting in abnormal gait patterns. While Ankle-Foot Orthoses (AFOs) are commonly used to address these issues, few are specifically designed for ankle varus. The Elastic Neutral Ankle-Foot Orthosis (EN-AFO) was developed [...] Read more.
Many stroke patients develop ankle deformities due to neurological or non-neurological factors, resulting in abnormal gait patterns. While Ankle-Foot Orthoses (AFOs) are commonly used to address these issues, few are specifically designed for ankle varus. The Elastic Neutral Ankle-Foot Orthosis (EN-AFO) was developed for this purpose. This study aimed to analyze changes in kinematic and kinetic gait data in stroke patients with ankle varus, comparing those walking with and without EN-AFO in both AFO and No-AFO groups. Initially, 30 stroke patients with ankle varus were screened; after exclusions, 17 were included in the final analysis. In the No-AFO group, EN-AFO significantly improved maximal ankle inversion on the affected side during the swing phase (from 4.63 ± 13.26 to 10.56 ± 11.40, p = 0.025). Similarly, in the AFO group, EN-AFO led to a significant improvement in maximal ankle inversion on the less-affected side during the swing phase (from 7.95 ± 10.11 to 12.01 ± 8.64, p = 0.021). Additionally, ground reaction forces on the affected side of the AFO group significantly increased at both the forefoot (from 182.76 ± 61.45 to 211.55 ± 70.57, p = 0.038) and hindfoot (from 210.67 ± 107.88 to 231.85 ± 105.38, p = 0.038) with EN-AFO. Conversely, maximal and minimal thoracic axial rotation on the affected side improved significantly in the No-AFO group compared to the AFO group with EN-AFO, during both the stance and swing phases (stance phase: max improvement from −1.13 ± 1.80 to 4.83 ± 8.05, min improvement from −1.06 ± 2.45 to 5.89 ± 7.56; swing phase: max improvement from −1.33 ± 2.13 to 5.49 ± 7.82, min improvement from −1.24 ± 2.43 to 5.95 ± 7.12; max p = 0.034, min p = 0.016 during stance; max p = 0.027, min p = 0.012 during swing). Furthermore, both maximal and minimal thoracic axial rotation on the less-affected side during the swing phase improved significantly in the No-AFO group (max improvement from −2.09 ± 4.18 to 6.04 ± 6.90, min improvement from −0.47 ± 2.13 to 8.18 ± 10.45; max p = 0.027, min p = 0.012) compared with the AFO group. These findings suggest that EN-AFO may effectively improve gait in stroke patients with ankle varus in the No-AFO group. Full article
Show Figures

Figure 1

Figure 1
<p>A flowchart of the study.</p>
Full article ">Figure 2
<p>The force plate (dark area) to analyze kinetic variables.</p>
Full article ">Figure 3
<p>The EN-AFO device in use. (a) Velcro straps; (b) fabric belt with elastic reinforcement; (c,d) elastic bands securing the lower leg and forefoot.</p>
Full article ">Figure 4
<p>The inner side of the EN-AFO. (a) Velcro strap; (b-1,b-2) joint between elastic band and fabric, (c) fabric, (d) the elastic support, (e) the elastic band wearing the lower leg and forefoot, (f-1,f-2) stiches in the shape of an overshoe in an area where thin plastic was attached. (Cited from Hwang and Park (2021) [<a href="#B10-actuators-13-00526" class="html-bibr">10</a>]).</p>
Full article ">Figure 5
<p>The attachments of the IMU sensors.</p>
Full article ">Figure 6
<p>Differentiation of kinematic data of the affected side in AFO and No-AFO groups after wearing Elastic Neutral AFO at stance phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Differentiation of kinematic data of the less-affected side in AFO and No-AFO groups after wearing Elastic Neutral AFO at stance phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Differentiation of kinematic data of the affected side after wearing Elastic Neutral AFO at swing phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Differentiation of kinematic data of the less-affected side after wearing Elastic Neutral AFO at swing phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Differentiation of kinematic data between AFO and NO-AFO groups after wearing EN-AFO at stance phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 11
<p>Differentiation of kinematic data between AFO and NO-AFO groups after wearing EN-AFO at swing phase (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 12
<p>Gait cycle in the AFO group: hip external rotation, ankle dorsiflexion, ankle inversion and foot external rotation.</p>
Full article ">Figure 13
<p>Gait cycle in the No-AFO group: hip external rotation, ankle dorsiflexion, ankle inversion, and foot external rotation.</p>
Full article ">
16 pages, 6286 KiB  
Article
The Biomechanical Effects of Kinesiology Taping Methods on Side-Step Cutting Movements in Chronic Ankle Instability
by Xuting Wang, Wenjing Quan, Yiwen Ma, Sarosi Jozsef, Yufei Fang and Yaodong Gu
Healthcare 2024, 12(24), 2561; https://doi.org/10.3390/healthcare12242561 - 19 Dec 2024
Viewed by 423
Abstract
Background: The ankle joint is among the most vulnerable areas for injuries during daily activities and sports. This study focuses on individuals with chronic ankle instability (CAI), comparing the biomechanical characteristics of the lower limb during side-step cutting under various conditions. The [...] Read more.
Background: The ankle joint is among the most vulnerable areas for injuries during daily activities and sports. This study focuses on individuals with chronic ankle instability (CAI), comparing the biomechanical characteristics of the lower limb during side-step cutting under various conditions. The aim is to analyze the impact of kinesiology tape (KT) length on the biomechanical properties of the lower limb during side-step cutting, thereby providing theoretical support and practical guidance for protective measures against lower-limb sports injuries. Methods: Twelve subjects with CAI who met the experimental criteria were recruited. Each subject underwent testing without taping (NT), with short kinesiology tape (ST), and with long kinesiology tape (LT), while performing a 45° side-step cutting task. This study employed the VICON three-dimensional motion capture system alongside the Kistler force plate to synchronously gather kinematic and kinetic data during the side-step cutting. Visual 3D software (V6.0, C-Motion, Germantown, MD, USA) was utilized to compute the kinematic and kinetic data, while OpenSim 4.4 software (Stanford University, Stanford, CA, USA) calculated joint forces. A one-way Analysis of Variance (ANOVA) was conducted using SnPM, with the significance threshold established at p < 0.05. The Origin software 2021 was used for data graphic processing. Results: KT was found to significantly affect joint angles, angular velocities, and moments in the sagittal, frontal, and transverse planes. LT increased hip and knee flexion angles as well as angular velocity, while ST resulted in reduced ankle inversion and increased knee internal rotation. Both types of KT enhanced hip abduction moment and knee adduction/abduction moment. Additionally, LT reduced the ankle joint reaction force. Conclusions: These findings suggest that the application of KT over a short duration leads to improvements in the lower-limb performance during side-step cutting motions in individuals with CAI, thus potentially decreasing the risk of injury. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Reflective marker’s front, side, and back position on subjects. (<b>b</b>) 45° side-step cutting experiment workflow.</p>
Full article ">Figure 2
<p>The short tape technique. (<b>a</b>) The first tape. (<b>b</b>) The second tape. (<b>c</b>) The third tape. (<b>d</b>) The fourth tape.</p>
Full article ">Figure 3
<p>The long tape technique. (<b>a</b>) The first tape. (<b>b</b>) The second tape. (<b>c</b>) The third tape. (<b>d</b>) The fourth tape.</p>
Full article ">Figure 4
<p>Workflow of data processing. (<b>a</b>) Experimental collection. (<b>b</b>) Static model. (<b>c</b>) The data processing of Visual 3D. (<b>d</b>) The results of hip, knee, and ankle joint moments. (<b>e</b>) Scaling of the model. (<b>f</b>) Inverse kinematics (IK). (<b>g</b>) Static optimization (SO). (<b>h</b>) Joint reaction analysis.</p>
Full article ">Figure 5
<p>The kinematic characteristics of the lower-limb joints during the side-step cutting stance phase. (<b>a</b>) The ankle joint in the frontal plane. (<b>b</b>) The hip joint in the sagittal plane. (<b>c</b>) The hip joint in the horizontal plane. (<b>d</b>) The knee joint in the sagittal plane. (<b>e</b>) The knee joint in the frontal plane. (<b>f</b>) The knee joint in the horizontal plane.</p>
Full article ">Figure 5 Cont.
<p>The kinematic characteristics of the lower-limb joints during the side-step cutting stance phase. (<b>a</b>) The ankle joint in the frontal plane. (<b>b</b>) The hip joint in the sagittal plane. (<b>c</b>) The hip joint in the horizontal plane. (<b>d</b>) The knee joint in the sagittal plane. (<b>e</b>) The knee joint in the frontal plane. (<b>f</b>) The knee joint in the horizontal plane.</p>
Full article ">Figure 6
<p>The lower-limb joints’ velocity in the sagittal, frontal, and transverse planes during the stance phase in side-step cutting. (<b>a</b>) The ankle joint in the horizontal plane. (<b>b</b>) The knee joint in the horizontal plane. (<b>c</b>) The hip joint in the sagittal plane. (<b>d</b>) The hip joint in the frontal plane. (<b>e</b>) The hip joint in the horizontal plane.</p>
Full article ">Figure 6 Cont.
<p>The lower-limb joints’ velocity in the sagittal, frontal, and transverse planes during the stance phase in side-step cutting. (<b>a</b>) The ankle joint in the horizontal plane. (<b>b</b>) The knee joint in the horizontal plane. (<b>c</b>) The hip joint in the sagittal plane. (<b>d</b>) The hip joint in the frontal plane. (<b>e</b>) The hip joint in the horizontal plane.</p>
Full article ">Figure 7
<p>The kinetics characteristics of the lower-limb moments during the side-step cutting stance phase. (<b>a</b>) The ankle joint in the frontal plane. (<b>b</b>) The ankle joint in the horizontal plane. (<b>c</b>) The knee joint in the frontal plane. (<b>d</b>) The knee joint in the horizontal plane. (<b>e</b>) The hip joint in the sagittal plane. (<b>f</b>) The hip joint in the frontal plane.</p>
Full article ">Figure 7 Cont.
<p>The kinetics characteristics of the lower-limb moments during the side-step cutting stance phase. (<b>a</b>) The ankle joint in the frontal plane. (<b>b</b>) The ankle joint in the horizontal plane. (<b>c</b>) The knee joint in the frontal plane. (<b>d</b>) The knee joint in the horizontal plane. (<b>e</b>) The hip joint in the sagittal plane. (<b>f</b>) The hip joint in the frontal plane.</p>
Full article ">Figure 8
<p>Joint reaction force characteristics during the support phase. (<b>a</b>) Hip joint reaction force. (<b>b</b>) Knee joint reaction force. (<b>c</b>) Ankle joint reaction force.</p>
Full article ">
19 pages, 26960 KiB  
Article
The Northern Giona Fault Zone, a Major Active Structure Through Central Greece
by Leonidas Gouliotis and Dimitrios Papanikolaou
GeoHazards 2024, 5(4), 1370-1388; https://doi.org/10.3390/geohazards5040065 - 18 Dec 2024
Viewed by 422
Abstract
The steep northern slopes of Giona Mt in central continental Greece are the result of an E-W normal fault dipping 35–45° to the north, extending from the Mornos River in the west to the village of Gravia in the east. This fault creates [...] Read more.
The steep northern slopes of Giona Mt in central continental Greece are the result of an E-W normal fault dipping 35–45° to the north, extending from the Mornos River in the west to the village of Gravia in the east. This fault creates a significant elevation difference of approximately 1500 m between the northern Giona footwall and the southern Iti hanging wall. The footwall comprises imbricated Mesozoic carbonates of the Parnassos unit, which exhibit large-scale drag folding near and parallel to the fault. The hanging wall comprises deformed sedimentary rocks of the Beotian unit and tectonic klippen of the Eastern Greece unit, forming a southward-tilted neotectonic block with subsidence near the Northern Giona Fault and uplift near the Ypati fault to the north. These two E-W faults represent younger structures disrupting the older NNW-trending tectonic framework. Fault scarps are observed all along the 14 km length of the Northern Giona fault accompanied by cataclastic zones, separating the carbonate formations of the Parnassos Unit from thick scree, slide blocks, boulders and olistholites. Inversion of fault-slip data has shown a mean slip vector of 45°, N004°E, which aligns with the current regional extensional deformation of the area, as confirmed by focal mechanism solutions. Based on the general asymmetry of the alpine units in the hanging wall, we interpret a listric fault geometry at depth using slip-line analysis and we forward modelled a disrupted fault-propagation fold using kinematic trishear algorithms, estimating a total displacement of 6500 m and a throw of approximately 2000 m. Seismic activity in the area of the Northern Giona Fault includes a magnitude 6.1 earthquake in 1852, which caused casualties, rockfalls and extensive damage, as well as a magnitude 5.1 event in 1983. The expected seismic magnitude is deterministically estimated between 6.2 and 6.7, depending on the potential westward continuation of the Northern Giona Fault beyond the Mornos River to the Northern Vardoussia saddle. The seismic hazard zone includes several villages located near the fault, particularly on the hanging wall, where intense landslide activity during seismic events could result in severe damage to regional infrastructure. The neotectonic development of the Northern Giona Fault highlights the importance of extending seismotectonic research into the mountainous regions of central Greece within the alpine formations, beyond the post-orogenic sedimentary basins. Full article
(This article belongs to the Special Issue Active Faulting and Seismicity—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Morphological map of the mountainous region of central Greece between the Corinth Gulf and the Sperchios Valley/Maliakos Gulf. The northern boundary is defined by the northern slopes of Iti Mt, where the Ypati Fault (YF) creates a significant topographic difference of 2000 m, separating the mountainous area from the Sperchios Valley. To the south, the northern slopes of Giona Mt. align with the Northern Giona Fault (NGF), marking a topographic difference of 1500 m between Giona Mt and Iti Mt.</p>
Full article ">Figure 2
<p>Three-dimensional perspective of the studied area with view from the east-northeast (<b>top</b>) and view from the west (<b>bottom</b>). In both views, the NGF and YF are indicated along the abrupt northern slopes of Giona and Iti Mts, respectively. These two subparallel faults have shaped the landscape, creating high mountain peaks at their northern edges on the uplifted footwalls and subsidence to the southern edges.</p>
Full article ">Figure 3
<p>Map illustrating the distribution of seismic epicenters for instrumental (solid circles) and historical earthquakes (purple triangles), along with focal mechanisms for M &gt; 4 (NOA—<a href="http://emsc-csem.org" target="_blank">http://emsc-csem.org</a>), GPS velocity vectors [<a href="#B7-geohazards-05-00065" class="html-bibr">7</a>] and neotectonic faults (black lines). Central Greece’s active deformation results primarily from displacements on E-W-trending faults and some NE-SW strike-slip events.</p>
Full article ">Figure 4
<p>(<b>Top</b>) Geotectonic map of the mountainous region of central Greece between the Sperchios valley to the north and the Corinth gulf to the south. Yellow dashed rectangle shows the extent of the detailed geological map along the NGF of Figure 6. (<b>Middle</b>) NNW—SSE cross section from the Orthrys Mt to the Corinth Gulf showing the geometry of the alpine units and the major neotectonic boundaries, including the YF and the NGF. Dashed blue line indicates the top carbonate of the Parnassos unit. (<b>Bottom</b>) 2D forward model across the NGF showing the deformation of a 10 km layer-cake model with the top horizon corresponding to the pre-Pliocene tectonic framework as built in Figure 11. The Mw 5.1 19 September 1983 earthquake is plotted on the profile alongside the NGF. 1: Pindos Unit, 2: Penteoria unit, 3: Vardoussia unit, 4: Parnassos unit, 5: Beotian unit, 6: Eastern Greece unit, 7: Late Oligocene–Miocene molassic sediments, 8: Late Miocene-Quaternary sediments, 9: Neotectonic and active fault, 10: Miocene Extensional Detachment, including the Itea-Amfissa detachment (IAD) 11: major thrust fault.</p>
Full article ">Figure 5
<p>Panoramic view looking SSW of the NGF (red dashed line) along the northern slopes of Giona Mt. The high-elevated area of northern Giona belonging to the Parnassos unit occur in the footwall whereas the ophiolites and related sediments of the uppermost Eastern Greece unit occur in the hanging wall.</p>
Full article ">Figure 6
<p>Geological map and cross section of the northern Giona region. 1: alluvial deposits, 2: scree deposits, 3: Neogene deposits, 4: Molassic sediments of Oligocene—Middle Miocene, 5: Triassic—Jurassic carbonates of the SubPelagonian unit, 6: Jurassic ophiolites, 7: Beotian Unit with Jurassic—Cretaceous pelagic limestones, 8: Eocene flysch of the Parnassos unit, 9: Mesozoic Carbonate platform of the Parnassos unit, with an older b2 (black) and a younger b3 (red) bauxite horizons in the cross section, 10: Eocene flysch of the Vardousssia unit, 11: Olistholites mainly of Carbonate rocks, 12: Neotectonic and active normal fault, 13: IAD, Itea-Amfissa Detachment, Miocene extensional Detachment, 14: Overthrust, 15: thrust, 16. Base of gravity slide. FW1, HW1: Imbricated tectonic units of the Parnassos nappe. FW, HW: NGF’s footwall and hanging wall.</p>
Full article ">Figure 7
<p>View to the west of the footwall at the central part of the NGF, directly east of the Vrayla peak, at 1800–2000 m altitude. The Parnassos carbonate sequence is characterized by two members in this site: a lower one of Late Jurassic age (Js) and an upper one of Late Jurassic-Early Cretaceous age (J13-K6) separated by a bauxite horizon (b2—[<a href="#B24-geohazards-05-00065" class="html-bibr">24</a>]). In the sketch, solid lines indicate W-dipping strata, while dashed lines indicate N-dipping strata. This change in dip azimuth is characteristic of a kilometric scale normal drag developed near the NGF.</p>
Full article ">Figure 8
<p>View from the east of the NGF. Two notable sites (<b>A</b>,<b>B</b>) where the grooved and striated fault surface is exposed and measurable, showing top-to-N movement.</p>
Full article ">Figure 9
<p>Outcrops of the fault surfaces along the northern Giona slopes. (<b>A</b>) Fault surface (red arrows) at the central part of the NGF. (<b>B</b>) Close view of a north-dipping fault plane along the high slopes at the western side of the NGF. (<b>C</b>) Characteristic cross-section of the fault zone in the eastern part of the NGF, showing a well-developed damage zone that grade to thick fault core (red arrow). (<b>D</b>) Curved fault surface at the eastern termination of the NGF close to the Gravia village.</p>
Full article ">Figure 10
<p>Fault-slip data and calculation of stress tensor with the method of direct inversion for the NGF. Lower hemisphere, equal-area stereographic projections. <b>Left pane</b> shows fault-slip data and the calculated stress axes (σ1, σ2, σ3). <b>Middle pane</b> is a fluctuation histogram of the deviation angle (angle between measured and calculated slip vectors) and stress ratio R(σ2 − σ3)/(σ1 − σ3). <b>Right pane</b> shows the P–T axes.</p>
Full article ">Figure 11
<p>Construction of a 10 km-layer cake model illustrating a five-stage progressive deformation of the NGF footwall and hanging wall through the application of a trishear kinematic model with increasing displacements of 1300 m, 2600 m, 3900 m, 5200 &amp; 6500 m. Details of the trishear model are provided within text. The top layer represents the regional pre-tectonic level, corresponding to the pre-Pliocene deformed state, which includes early orogenic Late Eocene thrust faults (solid lines with triangles) overlying the Parnassos flysch. The Parnassos flysch comprises two members: the lower red pelites and the upper pelitic-sandstone, separated by a dashed line. Unconformably overlying these units are late-orogenic Miocene molasse deposits (wavy brown line) within the Iti and northern Giona fault blocks. The northward-dipping listric geometry of the NGF at depth is based on slip-line analysis [<a href="#B45-geohazards-05-00065" class="html-bibr">45</a>].</p>
Full article ">Figure 12
<p>Map showing the spatial distribution of seismic intensity recorded by various catastrophic phenomena associated with the two significant earthquakes in the region. The solid ellipse represents the macroseismic intensities from the 14 July 1852 earthquake, which had a magnitude of 6.1 (yellow star). The dashed ellipse outlines the area affected by the microseismicity (magnitudes between 2.0 and 4.2) that followed the 19 September 1983, earthquake, which had a magnitude of 5.1 and a fault plane solution of an ENE-WSW normal fault.</p>
Full article ">
13 pages, 4785 KiB  
Article
SK-PSO: A Particle Swarm Optimization Framework with SOM and K-Means for Inverse Kinematics of Manipulators
by Fei Liu, Changqin Gao and Lisha Liu
Symmetry 2024, 16(12), 1667; https://doi.org/10.3390/sym16121667 - 17 Dec 2024
Viewed by 474
Abstract
In this paper, a particle swarm optimizer that integrates self-organizing maps and k-means clustering (SK-PSO) is proposed. This optimizer generates an asymmetric Cartesian space from random joint configurations when addressing the inverse kinematics of manipulators, followed by K-means clustering applied to the Cartesian [...] Read more.
In this paper, a particle swarm optimizer that integrates self-organizing maps and k-means clustering (SK-PSO) is proposed. This optimizer generates an asymmetric Cartesian space from random joint configurations when addressing the inverse kinematics of manipulators, followed by K-means clustering applied to the Cartesian space. The resulting clusters are used to reduce the dimensionality of the corresponding joint space using Self-Organizing Maps (SOM). During the solving process, the target point’s clustering region is determined first, and then the joint space point closest to the target point is selected as the initial population for the particle swarm algorithm. The simulation results demonstrate the effectiveness of the SK-PSO algorithm. Given the inherent asymmetry among different algorithms in handling the problem, SK-PSO achieves an average fitness value that is 0.02–0.62 times better than five other algorithms, with an asymmetric solving time that is only 0.03–0.34 times that of the other algorithms. Therefore, compared to the other algorithms, the SK-PSO algorithm offers high accuracy, speed, and precision. Full article
(This article belongs to the Section Engineering and Materials)
Show Figures

Figure 1

Figure 1
<p>Illustration of PSO for solving inverse kinematics.</p>
Full article ">Figure 2
<p>Illustration of SK-PSO for solving inverse kinematics.</p>
Full article ">Figure 3
<p>The step-by-step solving process of SK-PSO.</p>
Full article ">Figure 4
<p>Structure diagram of the UR5 manipulator.</p>
Full article ">Figure 5
<p>Initial posture of the manipulator.</p>
Full article ">Figure 6
<p>Desired posture of the manipulator.</p>
Full article ">Figure 7
<p>Iterative process of the SK-PSO and compared algorithms.</p>
Full article ">Figure 8
<p>Solution time of the SK-PSO and compared algorithms.</p>
Full article ">Figure 9
<p>Fifty points randomly within the workspace of the UR5 manipulator.</p>
Full article ">Figure 10
<p>Optimal result by SK-PSO and other compared algorithms.</p>
Full article ">Figure 11
<p>Solution time of SK-PSO and other compared algorithms.</p>
Full article ">
15 pages, 1709 KiB  
Article
An Aircraft-Manipulator System for Virtual Flight Testing of Longitudinal Flight Dynamics
by Ademayowa A. Ishola, James F. Whidborne and Gilbert Tang
Robotics 2024, 13(12), 179; https://doi.org/10.3390/robotics13120179 - 15 Dec 2024
Viewed by 531
Abstract
A virtual flight test is the process of flying an aircraft model inside a wind tunnel in a manner that replicates free-flight. In this paper, a 3-DOF aircraft-manipulator system is proposed that can be used for longitudinal dynamics virtual flight tests. The system [...] Read more.
A virtual flight test is the process of flying an aircraft model inside a wind tunnel in a manner that replicates free-flight. In this paper, a 3-DOF aircraft-manipulator system is proposed that can be used for longitudinal dynamics virtual flight tests. The system consists of a two rotational degrees-of-freedom manipulator arm with an aircraft wind tunnel model attached to the third joint. This aircraft-manipulator system is constrained to operate for only the longitudinal motion of the aircraft. Thus, the manipulator controls the surge and heave of the aircraft whilst the pitch is free to rotate and can be actively controlled by means of an all-moving tailplane of the aircraft if required. In this initial study, a flight dynamics model of the aircraft is used to obtain dynamic response trajectories of the aircraft in free-flight. A model of the coupled aircraft-manipulator system developed using the Euler method is presented, and PID controllers are used to control the manipulator so that the aircraft follows the free-flight trajectory (with respect to the air). The inverse kinematics are used to produce the reference joint angles for the manipulator. The system is simulated in MATLAB/Simulink and a virtual flight test trajectory is compared with a free-flight test trajectory, demonstrating the potential of the proposed system for virtual flight tests. Full article
(This article belongs to the Special Issue Adaptive and Nonlinear Control of Robotics)
Show Figures

Figure 1

Figure 1
<p>Schematic of the aircraft-manipulator system sited in the working section of an open-section wind tunnel (not to scale). The wind tunnel airspeed is shown as <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>W</mi> <mi>T</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>A 1/12th scaled BAe Hawk aircraft.</p>
Full article ">Figure 3
<p>Configuration for 3-DOF aircraft longitudinal dynamics.</p>
Full article ">Figure 4
<p>Manipulator aircraft system configuration.</p>
Full article ">Figure 5
<p>Air velocity as the sum of wind tunnel velocity and aircraft velocity.</p>
Full article ">Figure 6
<p>Block diagram showing system control architecture for the <span class="html-italic">i</span>th loop. The block denoted as IK represents the inverse kinematics’ reference source.</p>
Full article ">Figure 7
<p>Short period mode pitch response of the free-flying Hawk model and the aircraft-manipulator system (AMS) model.</p>
Full article ">Figure 8
<p>Phugoid mode pitch response of the free-flying Hawk model and the aircraft-manipulator system (AMS) model.</p>
Full article ">Figure 9
<p>Short period mode altitude response of the free-flying Hawk model and the aircraft-manipulator system (AMS) model.</p>
Full article ">Figure 10
<p>Phugoid mode height and lateral surge position response of the free-flying Hawk model (resolved into the wind tunnel axes) and the aircraft-manipulator system (AMS) model.</p>
Full article ">Figure 11
<p>Phugoid mode position response of the aircraft-manipulator system (AMS) model in the wind tunnel.</p>
Full article ">
21 pages, 19119 KiB  
Article
Caterpillar-Inspired Multi-Gait Generation Method for Series-Parallel Hybrid Segmented Robot
by Mingyuan Dou, Ning He, Jianhua Yang, Lile He, Jiaxuan Chen and Yaojiumin Zhang
Biomimetics 2024, 9(12), 754; https://doi.org/10.3390/biomimetics9120754 - 11 Dec 2024
Viewed by 565
Abstract
The body structures and motion stability of worm-like and snake-like robots have garnered significant research interest. Recently, innovative serial–parallel hybrid segmented robots have emerged as a fundamental platform for a wide range of motion modes. To address the hyper-redundancy characteristics of these hybrid [...] Read more.
The body structures and motion stability of worm-like and snake-like robots have garnered significant research interest. Recently, innovative serial–parallel hybrid segmented robots have emerged as a fundamental platform for a wide range of motion modes. To address the hyper-redundancy characteristics of these hybrid structures, we propose a novel caterpillar-inspired Stable Segment Update (SSU) gait generation approach, establishing a unified framework for multi-segment robot gait generation. Drawing inspiration from the locomotion of natural caterpillars, the segments are modeled as rigid bodies with six degrees of freedom (DOF). The SSU gait generation method is specifically designed to parameterize caterpillar-like gaits. An inverse kinematics solution is derived by analyzing the forward kinematics and identifying the minimum lifting segment, framing the problem as a single-segment end-effector tracking task. Three distinct parameter sets are introduced within the SSU method to account for the stability of robot motion. These parameters, represented as discrete hump waves, are intended to improve motion efficiency during locomotion. Furthermore, the trajectories for each swinging segment are determined through kinematic analysis. Experimental results validate the effectiveness of the proposed SSU multi-gait generation method, demonstrating the successful traversal of gaps and rough terrain. Full article
(This article belongs to the Section Locomotion and Bioinspired Robotics)
Show Figures

Figure 1

Figure 1
<p>Natural caterpillar locomotion pattern. (<b>a</b>) Natural caterpillar locomotion sequence (the red dashed line represents stable segment; the yellow dashed line represents swinging segment). (<b>b</b>) Schematic diagram of natural caterpillar segments.</p>
Full article ">Figure 2
<p>Nine-state of one segment motion trajectory.</p>
Full article ">Figure 3
<p>The hump formed on the natural caterpillar locomotion in a single segment 9-state trajectory. The illustration of the hump formed in the SSU method (red segment ((<b>3</b>)–(<b>6</b>) left) is the segment that is about to enter the swinging phase during the stance phase; red segment (right) is the segment that has ended the swinging phase during the stance phase. The yellow segment is the swinging segment in the swinging phase).</p>
Full article ">Figure 4
<p>Footfall-pattern diagram of nature caterpillar gait.</p>
Full article ">Figure 5
<p>Robot mechanism and variables. (<b>a</b>) 3-RSR. (<b>b</b>) 4-3-RSR.</p>
Full article ">Figure 6
<p>The kinematics analysis of 4-3-RSR robot SSU parameters <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>min</mi> </mrow> </msub> </mrow> </semantics></math>. In the 3-RSR parallel mechanism, (<b>a</b>) the relationship of the distal plate center in axis <math display="inline"><semantics> <mi>X</mi> </semantics></math> coordinate component and base angle <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> </mrow> </semantics></math>; (<b>b</b>) the relationship of the pitch angle of the distal plate and base angle <math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>θ</mi> <mn>3</mn> </msub> </mrow> </semantics></math>; (<b>c</b>) the relationship of the pitch angle of the distal plate and distal plate center in axis <math display="inline"><semantics> <mi>X</mi> </semantics></math> coordinate component. (<b>d</b>) The 2-3-RSR mechanism and variables.</p>
Full article ">Figure 7
<p>The robot posture when <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>SSU gait generation flowchart.</p>
Full article ">Figure 9
<p>The SSU swinging segment trajectory. (<b>a</b>) The gaits sequence for <math display="inline"><semantics> <mrow> <mfenced> <mrow> <mi>m</mi> <mo>−</mo> <mn>1</mn> </mrow> </mfenced> <mi>th</mi> </mrow> </semantics></math> segment trajectory <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>s</mi> <mo stretchy="false">^</mo> </mover> <mrow> <msub> <mi>c</mi> <mn>1</mn> </msub> </mrow> </msub> <mfenced> <mi>t</mi> </mfenced> </mrow> </semantics></math>. (<b>b</b>) The gaits sequence for <math display="inline"><semantics> <mrow> <mfenced> <mrow> <mi>m</mi> <mo>−</mo> <mn>1</mn> </mrow> </mfenced> <mi>th</mi> </mrow> </semantics></math> segment trajectory <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>s</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>c</mi> <mn>2</mn> </mrow> </msub> <mfenced> <mi>t</mi> </mfenced> </mrow> </semantics></math>. (<b>c</b>) The trajectory of <math display="inline"><semantics> <mrow> <mfenced> <mrow> <mi>m</mi> <mo>−</mo> <mn>2</mn> </mrow> </mfenced> <mi>th</mi> </mrow> </semantics></math> segment progressive. (<b>d</b>) The compensate trajectory <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>s</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>c</mi> <mn>1</mn> </mrow> </msub> <mfenced> <mi>t</mi> </mfenced> </mrow> </semantics></math> of <math display="inline"><semantics> <mrow> <mfenced> <mrow> <mi>m</mi> <mo>−</mo> <mn>1</mn> </mrow> </mfenced> <mi>th</mi> </mrow> </semantics></math> segment. (<b>e</b>) The compensate trajectory <math display="inline"><semantics> <mrow> <msub> <mover accent="true"> <mi>s</mi> <mo stretchy="false">^</mo> </mover> <mrow> <mi>c</mi> <mn>2</mn> </mrow> </msub> <mfenced> <mi>t</mi> </mfenced> </mrow> </semantics></math> of <math display="inline"><semantics> <mrow> <mfenced> <mrow> <mi>m</mi> <mo>−</mo> <mn>1</mn> </mrow> </mfenced> <mi>th</mi> </mrow> </semantics></math> segment progressive.</p>
Full article ">Figure 10
<p>Three gaits pattern of 4-3-RSR robot. (<b>a</b>) The 1-1-1-1-1 gait, (<b>b</b>) 1-1-2-1 gait, and (<b>c</b>) 1-2-2 gait. Footfall-pattern diagram of the (<b>d</b>) 1-1-1-1-1 gait, (<b>e</b>) 1-1-2-1 gait, and (<b>f</b>) 1-2-2 gait.</p>
Full article ">Figure 11
<p>The 4-3-RSR robot. (<b>a</b>) Three rotary joints replace the sphere joint. (<b>b</b>) The 4-3-RSR robot press plate (left) and main view (right).</p>
Full article ">Figure 12
<p>Joint trajectories of 4-3-RSR robot. (<b>a</b>) The 1-1-1-1-1 gait, (<b>b</b>) 1-1-2-1 gait, and (<b>c</b>) 1-2-2 gait, where (1) (2) (3) (4) illuminate the 1st, 2nd, 3rd, and 4th 3-RSR parallel mechanism joint trajectories.</p>
Full article ">Figure 13
<p>Three gaits experiment of the 4-3-RSR robot. (<b>a</b>) The 1-1-1-1-1 gait, (<b>b</b>) 1-1-2-1 gait, and (<b>c</b>) 1-2-2 gait. (The red dotted line represents the stable segment, and the yellow dotted line represents the swinging segment).</p>
Full article ">Figure 14
<p>Locomotion of the 4-3-RSR robot rectilinear gait.</p>
Full article ">Figure 15
<p>The 1-1-1-1-1-1 gait crossing gaps.</p>
Full article ">Figure 16
<p>The 1-1-1-1-1-1 gait on roughness terrain.</p>
Full article ">
23 pages, 3192 KiB  
Article
Design Optimization of a Parallel–Serial Manipulator Considering Stiffness Criteria
by Anton Antonov
Robotics 2024, 13(12), 176; https://doi.org/10.3390/robotics13120176 - 10 Dec 2024
Viewed by 654
Abstract
In this paper, we analyze stiffness and perform geometrical optimization of a parallel–serial manipulator with five degrees of freedom (5-DOF). The manipulator includes a 3-DOF redundantly actuated planar parallel mechanism, whose stiffness determines the stiffness of the whole mechanical system. First, we establish [...] Read more.
In this paper, we analyze stiffness and perform geometrical optimization of a parallel–serial manipulator with five degrees of freedom (5-DOF). The manipulator includes a 3-DOF redundantly actuated planar parallel mechanism, whose stiffness determines the stiffness of the whole mechanical system. First, we establish the kinematic and stiffness models of the mechanism and define its stiffness matrix. Two components of this matrix and the inverse of its condition number are chosen as stiffness indices. Next, we introduce an original two-step procedure for workspace analysis. In the first step, the chord method is used to find the workspace boundary. In the second step, the workspace is sampled inside the boundary by solving the point-in-polygon problem. After that, we derive stiffness maps and compute the average stiffness indices for various combinations of design variables. The number of these variables is reduced to two geometrical parameters, simplifying the representation and interpretation of the obtained results. Finally, we formulate the multi-objective design optimization problem, whose main goal is to maximize the lateral stiffness of the mechanism. We solve this problem using a hierarchical (ε-constraint) method. As a result, the lateral stiffness with optimized geometrical parameters increases by 54.1% compared with the initial design. Full article
Show Figures

Figure 1

Figure 1
<p>Organization of this paper.</p>
Full article ">Figure 2
<p>Considered manipulator: (<b>a</b>) computer model; (<b>b</b>) prototype during the operation.</p>
Full article ">Figure 3
<p>The parallel mechanism: (<b>a</b>) general configuration; (<b>b</b>) lowest configuration.</p>
Full article ">Figure 4
<p>Kinematic parameters of the mechanism.</p>
Full article ">Figure 5
<p>Determining the starting point on the workspace boundary.</p>
Full article ">Figure 6
<p>Computing the workspace boundary using the chord method: (<b>a</b>) finding point <math display="inline"><semantics> <msub> <mi>U</mi> <mrow> <mi>j</mi> <mo>+</mo> <mn>1</mn> </mrow> </msub> </semantics></math> by a circular search at point <math display="inline"><semantics> <msub> <mi>U</mi> <mi>j</mi> </msub> </semantics></math>; (<b>b</b>) angle <math display="inline"><semantics> <mi>ψ</mi> </semantics></math> depending on the <math display="inline"><semantics> <mi>σ</mi> </semantics></math> value.</p>
Full article ">Figure 7
<p>Sampling rectangle <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mn>1</mn> </msub> <msub> <mi>V</mi> <mn>2</mn> </msub> <msub> <mi>V</mi> <mn>3</mn> </msub> <msub> <mi>V</mi> <mn>4</mn> </msub> </mrow> </semantics></math> that envelops workspace boundary <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mn>1</mn> </msub> <msub> <mi>U</mi> <mn>2</mn> </msub> <mo>…</mo> <msub> <mi>U</mi> <mi>n</mi> </msub> </mrow> </semantics></math>. The blue and red dots are examples of samples inside and outside the boundary.</p>
Full article ">Figure 8
<p>Numbering of polygon vertices and edges: (<b>a</b>) the numbers near the vertices inside the parentheses correspond to values <math display="inline"><semantics> <msub> <mi>a</mi> <mi>j</mi> </msub> </semantics></math>, while the blue numbers near the edges are values <math display="inline"><semantics> <msub> <mi>b</mi> <mi>j</mi> </msub> </semantics></math>; (<b>b</b>) edge <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mn>1</mn> </msub> <msub> <mi>U</mi> <mn>2</mn> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>(</mo> <mn>1</mn> <mo>)</mo> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math> and edge <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mn>3</mn> </msub> <msub> <mi>U</mi> <mn>4</mn> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>b</mi> <mn>3</mn> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>(</mo> <mn>3</mn> <mo>)</mo> <mo>&lt;</mo> <mn>0</mn> </mrow> </semantics></math> are not counted in Algorithm 1, as they do not match the if statements of the algorithm.</p>
Full article ">Figure 9
<p>Workspace boundary for various end-effector orientations (<math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>10</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>20</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>30</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. The squares correspond to the boundary points.</p>
Full article ">Figure 10
<p>Workspace sampling for various parameters (<math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>325</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>425</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>345</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>455</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>Stiffness maps of <math display="inline"><semantics> <msub> <mi>k</mi> <mi>y</mi> </msub> </semantics></math> for various parameters (<math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>325</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>425</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>345</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>455</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>Stiffness maps of <math display="inline"><semantics> <msub> <mi>k</mi> <mi>z</mi> </msub> </semantics></math> for various parameters (<math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>325</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>425</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>345</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>455</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Stiffness map of <math display="inline"><semantics> <msup> <mi>k</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math> for various parameters (<math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>325</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>425</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>345</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>455</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>Stiffness maps for the case when the adjacent branches coincide (<math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; <math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) lateral stiffness <math display="inline"><semantics> <msub> <mi>k</mi> <mi>y</mi> </msub> </semantics></math>; (<b>b</b>) vertical stiffness <math display="inline"><semantics> <msub> <mi>k</mi> <mi>z</mi> </msub> </semantics></math>; (<b>c</b>) overall stiffness <math display="inline"><semantics> <msup> <mi>k</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>.</p>
Full article ">Figure 15
<p>The values of the performance metrics for various parameters <math display="inline"><semantics> <msub> <mi>y</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>y</mi> <mn>2</mn> </msub> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>φ</mi> <mo>=</mo> <msup> <mn>0</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>): (<b>a</b>) average lateral stiffness <math display="inline"><semantics> <msub> <mover accent="true"> <mi>k</mi> <mo stretchy="false">¯</mo> </mover> <mi>y</mi> </msub> </semantics></math>; (<b>b</b>) average vertical stiffness <math display="inline"><semantics> <msub> <mover accent="true"> <mi>k</mi> <mo stretchy="false">¯</mo> </mover> <mi>z</mi> </msub> </semantics></math>; (<b>c</b>) average overall stiffness <math display="inline"><semantics> <msup> <mover accent="true"> <mi>k</mi> <mo stretchy="false">¯</mo> </mover> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>; (<b>d</b>) workspace area <span class="html-italic">W</span>. The white circles correspond to the initial design with <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>255</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>505</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>; the white squares correspond to the optimized design with <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>345</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>y</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>575</mn> <mspace width="0.166667em"/> <mi>mm</mi> </mrow> </semantics></math>.</p>
Full article ">
31 pages, 8127 KiB  
Article
Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot
by Josué Goméz-Casas, Carlos A. Toro-Arcila, Nelly Abigaíl Rodríguez-Rosales, Jonathan Obregón-Flores, Daniela E. Ortíz-Ramos, Jesús Fernando Martínez-Villafañe and Oziel Gómez-Casas
Processes 2024, 12(12), 2831; https://doi.org/10.3390/pr12122831 - 10 Dec 2024
Viewed by 575
Abstract
This paper presents a data-driven kinematic model for the end-effector pose control applied to a variety of manipulator robots, focusing on the entire end-effector’s pose (position and orientation). The measured signals of the full pose and their computed derivatives, along with a linear [...] Read more.
This paper presents a data-driven kinematic model for the end-effector pose control applied to a variety of manipulator robots, focusing on the entire end-effector’s pose (position and orientation). The measured signals of the full pose and their computed derivatives, along with a linear combination of an estimated Jacobian matrix and a vector of joint velocities, generate a model estimation error. The Jacobian matrix is estimated using the Pseudo Jacobian Matrix (PJM) algorithm, which requires tuning only the step and weight parameters that scale the convergence of the model estimation error. The proposed control law is derived in two stages: the first one is part of an objective function minimization, and the second one is a constraint in a quasi-Lagrangian function. The control design parameters guarantee the control error convergence in a closed-loop configuration with adaptive behavior in terms of the dynamics of the estimated Jacobian matrix. The novelty of the approach lies in its ability to achieve superior tracking performance across different manipulator robots, validated through simulations. Quantitative results show that, compared to a classical inverse-kinematics approach, the proposed method achieves rapid convergence of performance indices (e.g., Root Mean Square Error (RMSE) reduced to near-zero in two cycles vs. a steady-state RMSE of 20 in the classical approach). Additionally, the proposed method minimizes joint drift, maintaining an RMSE of approximately 0.3 compared to 1.5 under the classical scheme. The control was validated by means of simulations featuring an UR5e manipulator with six Degrees of Freedom (DOF), a KUKA Youbot with eight DOF, and a KUKA Youbot Dual with thirteen DOF. The stability analysis of the closed-loop controller is demonstrated by means of the Lyapunov stability conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Closed-loop configuration diagram.</p>
Full article ">Figure 2
<p>CoppeliaSim simulation environment showcasing different types of robots. The figure highlights the reference frames of the world’s origin, end-effector, target object, and the orientation control joystick in each setup: (<b>a</b>) UR5 simulation setup, (<b>b</b>) KY simulation setup, and (<b>c</b>) KYD simulation setup.</p>
Full article ">Figure 3
<p>End-effector user commands.</p>
Full article ">Figure 4
<p>End-effector pose errors.</p>
Full article ">Figure 5
<p>End-effector velocities.</p>
Full article ">Figure 6
<p>Control signals.</p>
Full article ">Figure 7
<p>End-effector user commands.</p>
Full article ">Figure 8
<p>End-effector pose errors.</p>
Full article ">Figure 9
<p>End-effector velocities.</p>
Full article ">Figure 10
<p>Control signals.</p>
Full article ">Figure 11
<p>Commands for the left tip.</p>
Full article ">Figure 12
<p>Commands for the right tip.</p>
Full article ">Figure 13
<p>Pose errors of the left tip.</p>
Full article ">Figure 14
<p>Pose errors of the right tip.</p>
Full article ">Figure 15
<p>Left end-effector velocities.</p>
Full article ">Figure 16
<p>Right end-effector velocities.</p>
Full article ">Figure 17
<p>Control signals.</p>
Full article ">Figure 18
<p>CoppeliaSim simulation environment: the KY’s end-effector performs a trajectory tracking task defined in the <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>x</mi> <mo>,</mo> <mi>y</mi> <mo>)</mo> </mrow> </semantics></math> plane, with a constant height <span class="html-italic">z</span>, while keeping the initial orientation unchanged—the goal is to perform a polishing operation under the surface of the disk. (<b>a</b>) Simulation setup home position. (<b>b</b>) KY under DDMC scheme. (<b>c</b>) KY under classic inverse-kinematics control scheme.</p>
Full article ">Figure 19
<p>Pose control performance indexes while performing a cyclic routine. (<b>a</b>) Performance indexes under DDCM scheme. (<b>b</b>) Performance indexes under classic inverse-kinematics control scheme.</p>
Full article ">
22 pages, 6936 KiB  
Article
Design and Performance Analysis of a Parallel Pipeline Robot
by Zhonghua Shen, Menglin Xie, Zhendong Song and Danyang Bao
Electronics 2024, 13(23), 4848; https://doi.org/10.3390/electronics13234848 - 9 Dec 2024
Viewed by 437
Abstract
A parallel four-legged pipeline robot is designed to mitigate the issue of uneven motor loading on the single-leg linkage responsible for movement along the pipe diameter. This issue occurs because the drive motor located closer to the robot body requires higher torque when [...] Read more.
A parallel four-legged pipeline robot is designed to mitigate the issue of uneven motor loading on the single-leg linkage responsible for movement along the pipe diameter. This issue occurs because the drive motor located closer to the robot body requires higher torque when the serial robot operates along the inner wall of a circular polyethylene gas pipe in an urban environment. The forward and inverse kinematic equations for a single-leg linkage are derived to establish the relationship between joint angles and foot trajectories. Building on this analysis, the forward and inverse kinematic solutions for all four legs are also derived. An optimized diagonal trotting gait is selected as the robot’s walking pattern to ensure a balance between stability and movement efficiency, considering the robot’s structural configuration. Motion simulations for both the serial and parallel robots are performed using simulation software, with a detailed analysis of the displacement of the robot’s center of mass and the leg centers during movement. The driving torque of the leg motors in both configurations is controlled and examined. Simulation results indicate that the designed parallel four-legged pipeline robot achieves lower motion error and smoother leg movements within the pipe. Compared to the serial robot, the maximum torque required to drive the leg motors is reduced by approximately 33%, demonstrating the effectiveness and validity of the overall structural design. Full article
Show Figures

Figure 1

Figure 1
<p>Overall display.</p>
Full article ">Figure 2
<p>Motion structure diagram of parallel robot 1. 1—Hip joint drive motor, 2—Hip joint plate, 3—Hip joint linkage, 4—Second mobile joint drive motor, 5—Second mobile joint active linkage, 6—Second mobile joint passive linkage, 7—Sucker, 8—Sucker connecting member, 9—First mobile joint passive linkage, 10—First mobile joint active linkage, 11—First mobile joint drive motor, 12—Body part.</p>
Full article ">Figure 3
<p>Motion structure diagram of parallel robot 2.</p>
Full article ">Figure 4
<p>Initial position of parallel pipeline robot.</p>
Full article ">Figure 5
<p>Coordinate system of robot leg single chain.</p>
Full article ">Figure 6
<p>Plan–coordinate system of leg single chain.</p>
Full article ">Figure 7
<p>Plan–coordinate system of leg single chain 2.</p>
Full article ">Figure 8
<p>Four different gaits.</p>
Full article ">Figure 9
<p>Axial gait diagram.</p>
Full article ">Figure 10
<p>Sequence chart.</p>
Full article ">Figure 11
<p>Surface mesh generation.</p>
Full article ">Figure 12
<p>Initial state of the tandem and the parallel.</p>
Full article ">Figure 13
<p>Both the parallel type’s and tandem type’s moving joint of the LB.</p>
Full article ">Figure 14
<p>Both the parallel type’s and tandem type’s moving joint of the LF.</p>
Full article ">Figure 15
<p>The centroid displacement of the robot’s body part.</p>
Full article ">Figure 16
<p>Vertical displacement of legs.</p>
Full article ">Figure 17
<p>Both the parallel type’s and tandem type’s hip joint of the LB.</p>
Full article ">Figure 18
<p>Both the parallel type’s and tandem type’s moving joint of the LF.</p>
Full article ">Figure 19
<p>Both the parallel type’s and tandem type’s moving joint of the LB.</p>
Full article ">Figure 20
<p>The centroid displacement of the robot’s body part.</p>
Full article ">
33 pages, 24137 KiB  
Article
Development of a Reduced-Degree-of-Freedom (DOF) Bipedal Robot with Elastic Ankles
by Sharafatdin Yessirkepov, Michele Folgheraiter, Arman Abakov and Timur Umurzakov
Robotics 2024, 13(12), 172; https://doi.org/10.3390/robotics13120172 - 4 Dec 2024
Viewed by 778
Abstract
One of the most challenging aspects of designing a humanoid robot is ensuring stable walking. To achieve this, the kinematic architecture must support 3D motion and maintain equilibrium, particularly during single-foot support. Without proper configuration, the robot may experience unbalanced weight distribution, significantly [...] Read more.
One of the most challenging aspects of designing a humanoid robot is ensuring stable walking. To achieve this, the kinematic architecture must support 3D motion and maintain equilibrium, particularly during single-foot support. Without proper configuration, the robot may experience unbalanced weight distribution, significantly increasing the risk of falling while walking. While adding redundant degrees of freedom (DOFs) can enhance adaptability, it also raises the system’s complexity and cost and the need for more sophisticated control strategies and higher energy consumption. This paper explores a reduced-DOF bipedal robot, which, despite its limited number of DOFs, is capable of performing 3D motion. It features an inverted pendulum and elastic ankles made of thermoplastic polyurethane (TPU), enabling effective balance control and attenuation of disturbances. The robot’s electromechanical design is introduced alongside the kinematic model. Momentum equilibrium in a pseudo-static mode is considered in both the frontal and sagittal planes, taking into account the pendulum and the swinging leg during the single support phase. The TPU ankle’s performance is assessed based on its ability to resist external bending forces, highlighting challenges related to the robot’s weight equilibrium stability and ankle inversion. Experimental results from both Finite Element Analysis (FEA) and real-world tests are compared. Lastly, the joint movements of the inverted pendulum-based biped robot are evaluated in both a virtual environment and a physical prototype while performing lateral tilting and various gait sequences. Full article
(This article belongs to the Section Intelligent Robots and Mechatronics)
Show Figures

Figure 1

Figure 1
<p>A novel 5-DOF biped robot kinematic architecture with a pendulum balancer: (<b>a</b>) CAD design developed by FreeCAD 0.19.3 software. (<b>b</b>) Real prototype. (<b>c</b>,<b>d</b>) Numerical model. 1—Weight applied by the end-effector frame of a pendulum; 2—Pendulum link; 3—Inverted pendulum actuator; 4—Central link; 5—Left hip roll servo motor; 6—Left leg link; 7—Left leg yaw servo motor. 8—IMU sensor (Model: WitMotion); 9—Dual motor controllers (Model: Odrive 3.5). <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>1</mn> </msub> </semantics></math>—Right leg yaw actuator’s position; <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math>—Right hip roll joint’s position; <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>3</mn> </msub> </semantics></math>—Left hip roll joint’s position; <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>4</mn> </msub> </semantics></math>—Left leg yaw actuator’s position; <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>6</mn> </msub> </semantics></math>—Upper limb pendulum actuator’s position.</p>
Full article ">Figure 2
<p>CAD design of a novel reduced-DOF bipedal robot with an RRYY kinematic architecture performing a half gait cycle: (<b>a</b>) Home position. (<b>b</b>) Rotation of the upper limb part to the right side. (<b>c</b>) Stance phase. (<b>d</b>) Forward movement. (<b>e</b>) Double support phase. (<b>f</b>) Returning the upper limb part to the home position. (<b>g</b>) Closeup view of the feet. Δx—Displacement of the robot.</p>
Full article ">Figure 3
<p>Description of the chains used in kinematic calculations for the RRYY bipedal robot: (<b>a</b>) first kinematic chain (frames 1, 2, 3, 4 and 5); (<b>b</b>) second kinematic chain (frames 1, 2, 6 and 7); (<b>c</b>) intersection between the first and second kinematic chains (frames 1 and 2). <math display="inline"><semantics> <msub> <mi mathvariant="normal">L</mi> <mi>i</mi> </msub> </semantics></math>—link length that corresponds to frame i, m.</p>
Full article ">Figure 4
<p>Sketch of the bipedal robot stepping forward while balancing on its right leg: (<b>a</b>) Frontal plane. (<b>b</b>) Sagittal plane. (<b>c</b>) Transverse plane.</p>
Full article ">Figure 5
<p>Sketch of a weight balancer with weights applied on both sides.</p>
Full article ">Figure 6
<p>TPU -based ankle inclination on the frontal plane and its geometric structure: (<b>a</b>) abnormal case; (<b>b</b>) normal case; (<b>c</b>) foot structure; (<b>d</b>) 3D view of an ankle; (<b>e</b>) 2D Surface view of an ankle. 1—IMU sensor; 2—aluminum extrusion 20 mm × 20 mm; 3—foot cover; 4—TPU-based ankle; 5—back part; 6—front part of a foot.</p>
Full article ">Figure 7
<p>Pictures of the 3D printed sample TPU ankle prototypes with a modulus elasticity of <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 90 MPa at various geometries a × b × h: (<b>a</b>) 10 mm × 10 mm × 20 mm; (<b>b</b>) 20 mm × 10 mm × 20 mm; (<b>c</b>) 10 mm × 10 mm × 40 mm; (<b>d</b>) 20 mm × 10 mm × 40 mm; (<b>e</b>) 20 mm × 20 mm × 40 mm.</p>
Full article ">Figure 8
<p>Pictures of the tested TPU ankles with a modulus elasticity of <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 90 MPa and link length of 500 mm at different geometries. Graphs of inclination angle <math display="inline"><semantics> <mi>α</mi> </semantics></math> (deg) vs. external tension force F (N): (<b>a</b>) FEA method; (<b>b</b>) Real prototypes. 1—IMU sensor. 2—Force sensor.</p>
Full article ">Figure 9
<p>Graphs of the tested TPU ankles with a modulus elasticity of <math display="inline"><semantics> <mi>σ</mi> </semantics></math> = 90 MPa and link length of 500 mm at different geometries with axes representing inclination angle <math display="inline"><semantics> <mi>α</mi> </semantics></math> (deg) vs. tension force F (N): (<b>a</b>) FEA method; (<b>b</b>) Real prototypes.</p>
Full article ">Figure 10
<p>Diagrams of estimated inclination angle minimization: (<b>a</b>) FEA method; (<b>b</b>) real prototypes; (<b>c</b>) average result.</p>
Full article ">Figure 11
<p>Block diagram dedicated to the RRYY biped robot single support phase stability control.</p>
Full article ">Figure 12
<p>Elastic ankle stability control graphs during the single support phase: (<b>a</b>) Central link positions measured with an IMU sensor. (<b>b</b>) Right hip roll joint angles measured with a position encoder.</p>
Full article ">Figure 13
<p>RRYY bipedal robot lifting the left leg by 45 degrees and putting it down at a single-support stance (developed in the CoppeliaSim environment).</p>
Full article ">Figure 14
<p>Kinematic results of RRYY bipedal robot’s actuators lifting the left leg by 45 degrees and putting down in the single-support stance: (<b>a</b>) Joint positions (virtual environment). (<b>b</b>) Joint velocities (virtual environment). (<b>c</b>) Joint positions (real prototype). (<b>d</b>) Joint velocities (real prototype).</p>
Full article ">Figure 15
<p>Real model of the RRYY bipedal robot lifting the left leg by 45 degrees and putting it down in the single support stance.</p>
Full article ">Figure 16
<p>RRYY bipedal robot performing lateral motion lasting 2 cycles (developed in CoppeliaSim environment).</p>
Full article ">Figure 17
<p>Graphs of biped robot’s joints were obtained during the execution of a lateral movement: (<b>a</b>) Joint positions (virtual environment). (<b>b</b>) Joint velocities (virtual environment). (<b>c</b>) Joint positions (real prototype). (<b>d</b>) Joint velocities (real prototype).</p>
Full article ">Figure 18
<p>Testing an RRYY bipedal robot based on a lateral motion with a duration of 2 cycles.</p>
Full article ">Figure 19
<p>Kinematic results of 5 joints were obtained during the execution of a gait sequence: (<b>a</b>) Joint positions (virtual environment). (<b>b</b>) Joint velocities (virtual environment). (<b>c</b>) Joint positions (real prototype). (<b>d</b>) Joint velocities (real prototype).</p>
Full article ">Figure 20
<p>Pseudo-static walking sequence with 2 cycles (CoppeliaSim environment).</p>
Full article ">Figure 21
<p>Pseudo-static walking sequence with 2 cycles (Physical prototype).</p>
Full article ">Figure 22
<p>Experimental results of 5 joints obtained from numerical and real models after the execution of two gait cycles: (<b>a</b>) Torque computed from the numerical model. (<b>b</b>) Torque measured. (<b>c</b>) Joint mechanical power. (<b>d</b>) Joint electrical power.</p>
Full article ">Figure A1
<p>Block diagram of electrical and software connections.</p>
Full article ">Figure A2
<p>Right triangles were used to determine the right hip roll angle <math display="inline"><semantics> <msub> <mi>θ</mi> <mn>2</mn> </msub> </semantics></math> during the inverse kinematics (IK) calculation: (<b>a</b>) Swinging leg part. (<b>b</b>) Pendulum actuator part.</p>
Full article ">
17 pages, 6358 KiB  
Article
Continuous Multi-Target Approaching Control of Hyper-Redundant Manipulators Based on Reinforcement Learning
by Han Xu, Chen Xue, Quan Chen, Jun Yang and Bin Liang
Mathematics 2024, 12(23), 3822; https://doi.org/10.3390/math12233822 - 3 Dec 2024
Viewed by 667
Abstract
Hyper-redundant manipulators based on bionic structures offer superior dexterity due to their large number of degrees of freedom (DOFs) and slim bodies. However, controlling these manipulators is challenging because of infinite inverse kinematic solutions. In this paper, we present a novel reinforcement learning-based [...] Read more.
Hyper-redundant manipulators based on bionic structures offer superior dexterity due to their large number of degrees of freedom (DOFs) and slim bodies. However, controlling these manipulators is challenging because of infinite inverse kinematic solutions. In this paper, we present a novel reinforcement learning-based control method for hyper-redundant manipulators, integrating path and configuration planning. First, we introduced a deep reinforcement learning-based control method for a multi-target approach, eliminating the need for complicated reward engineering. Then, we optimized the network structure and joint space target points sampling to implement precise control. Furthermore, we designed a variable-reset cycle technique for a continuous multi-target approach without resetting the manipulator, enabling it to complete end-effector trajectory tracking tasks. Finally, we verified the proposed control method in a dynamic simulation environment. The results demonstrate the effectiveness of our approach, achieving a success rate of 98.32% with a 134% improvement using the variable-reset cycle technique. Full article
(This article belongs to the Special Issue Intelligent Control and Applications of Nonlinear Dynamic System)
Show Figures

Figure 1

Figure 1
<p>Overall model of the hyper-redundant manipulator.</p>
Full article ">Figure 2
<p>D-H coordinate frame of the hyper-redundant manipulator.</p>
Full article ">Figure 3
<p>Diagram of reinforcement learning process of this paper.</p>
Full article ">Figure 4
<p>Structure of the critic networks and actor networks. (<b>a</b>) DenseConnect. (<b>b</b>) SimpleDenseConnect.</p>
Full article ">Figure 5
<p>Distribution of the target points mapped to a workspace by different sampling in the joint space of a 12-link, 24-DOF, and hyper-redundant manipulator. (<b>a</b>) Normal Distribution I sampling. (<b>b</b>) Normal Distribution II sampling. (<b>c</b>) Uniform distribution sampling. (<b>d</b>) U-shaped distribution sampling.</p>
Full article ">Figure 6
<p>Different control tasks of the hyper-redundant manipulator. (<b>a</b>) Conventional multi-target approach control. (<b>b</b>) Continuous multi-target approach control.</p>
Full article ">Figure 7
<p>Problems of the direct samples from joint space. (<b>a</b>) Initial state model penetration. (<b>b</b>) Knotting. (<b>c</b>) Collision. (<b>d</b>) Inconsistent with reality.</p>
Full article ">Figure 8
<p>Performance of the 12-link, 24-DOF, and hyper-redundant manipulator on continuous multi-target approaching tasks trained by our method. (<b>a</b>) Circle on <math display="inline"><semantics> <mrow> <mi>Y</mi> <mi>O</mi> <mi>Z</mi> </mrow> </semantics></math> plane. (<b>b</b>) Circle on <math display="inline"><semantics> <mrow> <mi>X</mi> <mi>O</mi> <mi>Y</mi> </mrow> </semantics></math> plane. (<b>c</b>) Custom shape on <math display="inline"><semantics> <mrow> <mi>Y</mi> <mi>O</mi> <mi>Z</mi> </mrow> </semantics></math> plane.</p>
Full article ">Figure 9
<p>Performance of the 12-link, 24-DOF, and hyper-redundant manipulator on continuous multi-target approaching tasks trained by each episode reset. (<b>a</b>) TD3. (<b>b</b>) PPO.</p>
Full article ">Figure 10
<p>Training results of the different network structures under different control accuracies.</p>
Full article ">Figure 11
<p>Performance of the 12-link, 24-DOF, and hyper-redundant manipulator on the continuous multi-target approaching tasks trained by each episode reset. (<b>a</b>) Circle on <math display="inline"><semantics> <mrow> <mi>Y</mi> <mi>O</mi> <mi>Z</mi> </mrow> </semantics></math> plane. (<b>b</b>) Circle on <math display="inline"><semantics> <mrow> <mi>X</mi> <mi>O</mi> <mi>Y</mi> </mrow> </semantics></math> plane.</p>
Full article ">Figure 12
<p>Training effect of the planar hyper-redundant manipulator with 2-to-12 DOFs that were trained by never resetting.</p>
Full article ">Figure 13
<p>Training effect of the hyper-redundant manipulator by fixed-cycle reset. (<b>a</b>) Planar manipulator with 10 DOFs. (<b>b</b>) 3D manipulators with 4-to-20 DOFs.</p>
Full article ">
Back to TopTop