[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
Next Article in Journal
Coffee-Waste-Based ZnCl2 Activated Carbon in High-Performance Supercapacitor Electrodes: Impact of Graphitization, Surface Morphology, Porosity and Conductivity
Previous Article in Journal
A New Productivity Evaluation Method for Horizontal Wells in Offshore Low-Permeability Reservoir Based on Modified Theoretical Model
Previous Article in Special Issue
Stochastic Plantwide Optimizing Control for an Acrylic Acid Plant
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot

by
Josué Goméz-Casas
1,
Carlos A. Toro-Arcila
1,*,
Nelly Abigaíl Rodríguez-Rosales
2,
Jonathan Obregón-Flores
1,
Daniela E. Ortíz-Ramos
1,
Jesús Fernando Martínez-Villafañe
1 and
Oziel Gómez-Casas
1
1
Faculty of Engineering, Universidad Autónoma de Coahuila, Fundadores Km. 13, Las Glorias, Ciudad Universitaria, Arteaga 25350, Coahuila, Mexico
2
Department of Metal Mechanics, Tecnológico Nacional de México/I.T. Saltillo, Boulevard Venustiano Carranza 2400, Colonia Tecnológico, Saltillo 25280, Coahuila, Mexico
*
Author to whom correspondence should be addressed.
Processes 2024, 12(12), 2831; https://doi.org/10.3390/pr12122831
Submission received: 1 November 2024 / Revised: 30 November 2024 / Accepted: 4 December 2024 / Published: 10 December 2024
Figure 1
<p>Closed-loop configuration diagram.</p> ">
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> ">
Figure 3
<p>End-effector user commands.</p> ">
Figure 4
<p>End-effector pose errors.</p> ">
Figure 5
<p>End-effector velocities.</p> ">
Figure 6
<p>Control signals.</p> ">
Figure 7
<p>End-effector user commands.</p> ">
Figure 8
<p>End-effector pose errors.</p> ">
Figure 9
<p>End-effector velocities.</p> ">
Figure 10
<p>Control signals.</p> ">
Figure 11
<p>Commands for the left tip.</p> ">
Figure 12
<p>Commands for the right tip.</p> ">
Figure 13
<p>Pose errors of the left tip.</p> ">
Figure 14
<p>Pose errors of the right tip.</p> ">
Figure 15
<p>Left end-effector velocities.</p> ">
Figure 16
<p>Right end-effector velocities.</p> ">
Figure 17
<p>Control signals.</p> ">
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> ">
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> ">
Versions Notes

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 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.

1. Introduction

In recent years, robotic systems have become integral to daily life and significantly improve the quality of human activities. According to the Executive Summary of World Robotics 2022, the International Federation of Robotics (IFR) currently recognizes 1010 service robot producers worldwide. These service robots are applied in various domains, including transportation, hospitality, professional cleaning, agriculture, and construction. Of these, manipulator robots have emerged as the primary choice for automation in industries due to their multidegree-of-freedom (DOF) capabilities, which allow them to perform automatic and reprogrammable tasks such as welding, painting, picking and placing, and polishing processes (Islam et al. [1], Yu et al. [2], Gracia et al. [3], Saxena et al. [4]).
A significant challenge in industrial automation is controlling the complete pose (position and orientation) of a robot’s end-effector in the task space during complex operations (Awan et al. [5]). Traditional kinematic approaches, which rely on precise models, perform well under ideal conditions but are often limited when such information is incomplete, subject to change, or unavailable. This work addresses these challenges by proposing a data-driven method for trajectory tracking and pose regulation in the task space. Unlike conventional approaches, the proposed method eliminates the need for prior knowledge of the physical parameters of the robot and avoids the reliance on encoder-based feedback. This makes it particularly advantageous in scenarios involving unmodeled dynamics, parameter uncertainties, or sensor limitations, such as operations in unstructured environments or applications requiring hardware simplification. Using data-driven control, this approach broadens the applicability of robotic manipulators to scenarios where traditional methods may struggle to deliver satisfactory performance.
During the last decade, Data-Driven Modeling and Control (DDMC) has emerged as a promising option to improve autonomy in industrial processes (Hou et al. [6]). DDMC utilizes online data transmission in closed-loop controlled systems that involve sensors, actuators, and processing units. When referring to efficiency in DDMC, it implies that fewer hardware components, such as encoders, are necessary. In this scenario, the system relies solely on the feedback signal from the end-effector pose. Consequently, the probability of failure is reduced because the system is dependent on a single measurement source rather than on multiple sources (encoder case). This data-driven approach has become a novel strategy for industrial processes that require control of the 3D Euclidean coordinates of the end-effector (task space), which traditionally relies on a completely parameterized analytical model (Treesatayapun [7]). Given that DDMC only requires the association of output signals with input signals of unknown robotic system, this approach is applicable to rigid, soft, inertial, and non-inertial robots (Treesatayapun [8], Campos-Torres et al. [9], Gao et al. [10], Liu and Hou [11]). For a robotic arm, the joint velocity serves as the input signal, while the end-effector velocity acts as the output signal (Gómez et al. [12]). Consequently, the relationship between the end-effector and joint velocities allows for the online approximation of a Jacobian matrix in a model-free approach. The Jacobian matrix is essential for the kinematic model of a robot, particularly for applying closed-loop control in the task space of the end-effector. Then, DDMC in robotic systems typically occurs in two stages: The first stage involves the online computation of the Jacobian matrix using an estimation algorithm. For example, in Li et al. [13], the Jacobian matrix was estimated to achieve trajectory tracking for a continuum robot using an adaptive Kalman filter algorithm, while in Gómez et al. [12] it was estimated for a redundant 8-DOF robot using the Pseudo Jacobian Matrix (PJM) algorithm. The second stage involves designing the control law in the task space, using the pose of the end-effector as feedback, with the inverted estimated Jacobian mapping control signals into joint coordinates. It is important to note that the closed-loop system assumes an equivalent inverse mapping between inputs and outputs; therefore, the model estimation error must converge before the control error in a closed-loop configuration, otherwise, tasks performed by the end-effector may fail.
Although geometric data for robotic manipulators are often easily accessible from manufacturers’ documentation or open-source libraries such as ROS (The Robot Operating System), these models may not always be suitable for real-time control, especially in environments where the robot operates under variable or uncertain conditions. In addition, model-free approaches are particularly advantageous when dealing with robots for which kinematic models may be imprecise, unavailable, or subject to change. Consequently, a data-driven method that estimates the manipulator’s Jacobian matrix in real time offers flexibility and robustness across a wide range of applications where traditional models may be inadequate.
Controlling the complete pose of a robot’s end-effector is challenging for traditional kinematic models, primarily due to the need to represent the end-effector’s orientation (Villalobos et al. [14], Abbes and Poisson [15]). The control of angular velocity must be expressed as a function of joint velocities and the change rates along the axes of a chosen frame, which is standard for orientation tasks (Campa and De La Torre [16]). Modifying how angular velocity is represented results in changes to the orientation partition of the Jacobian matrix if it is computed. Common body orientation representations include Euler angles, which are intuitive but suffer from computational issues (Bai et al. [17]), and unit quaternions, which, while less intuitive, provide better numerical stability (Gonçalves et al. [18]). Other methods for representing the orientation of rigid bodies include rotation matrices, angle-axis pairs, and Euler parameters, each having its own strengths and weaknesses, including the presence of singularities. In the last decade, the DDMC has emerged as an attractive alternative from the control point of view, considering that it only requires the online input and output signals to estimate an equivalent model of the controlled plant; moreover, this approach dispenses with online input to compute an offline exhaustive analytical model. The data-driven approach applications were validated to Single-Input and Single Output (SISO) systems (Treesatayapun [8]). However, recently, the data-driven approach has been extended to MIMO robotic systems, increasing the complexity in this approach; hence, the estimation of the Jacobian matrix relies on the quality of the input and output signals and the number of robot DOFs. Some works have presented the position control of the robot end-effector (Gómez et al. [19]). Therefore, the challenge in this article is to extend the difficulty to the entire end-effector pose control (position and orientation). The innovative proposed algorithm supports all angle representations, consequently alleviating singularity-related challenges that typically complicate orientation control algorithms.
Existing approaches that address similar challenges for manipulator robots primarily focus on the end-effector position rather than on the complete pose. Singularities in orientation representations can complicate traditional orientation control methods. In the method proposed in this work, any choice of orientation representation is compatible with the PJM algorithm. Another common issue with full-pose control is that some robots, particularly those with fixed base frames, may struggle to achieve certain desired position and orientation coordinates due to mobility constraints.
In Nguyen et al. [20], an iterative learning algorithm was introduced for functional approximation, demonstrated through the use of a Radial Basis Function (RBF) neural network to approximate the kinematic models of two distinct robots, an SCARA with four DOF and an UR10 with six DOF, focusing solely on the position of the end-effector. Additionally, in Nguyen and Cheah [21], a data-driven technique was proposed employing a Neural Network (NN) to approximate robot kinematics without prior knowledge of the robot’s structure. Their method’s efficacy was showcased through simulation results on a forward kinematics task for an UR5e robot (six DOF) using an offline simulator, considering only the end-effector position in the task space. Furthermore, in Wu et al. [22], a model-free adaptive predictive tracking control approach was presented for the PUMA560 robotic manipulator (six DOF) with uncertain parameters. The effectiveness of the algorithm was validated through numerical simulations focusing on tracking circular trajectories while considering only the end-effector’s position. In Fan et al. [23], a kinematics-focused data-driven motion force control scheme was proposed, using a seven-DOF robot manipulator. They considered forces only along the z-axis, with robot movement confined to the x–y plane and a constant orientation of the end-effector. Moreover, in Marconi et al. [24], a structured prediction strategy was introduced to learn the inverse kinematics of redundant robots, blending data-driven techniques with forward kinematics. They empirically demonstrated that their method can compensate for forward kinematic bias and achieve accurate inverse kinematics, particularly when the available kinematic model is imprecise. Their approach is aimed at estimating the joint configuration needed to attain the desired pose, evaluated through trajectory reconstruction tasks. However, they noted orientation challenges compared to selectively damped least squares and encountered execution delays during inference, hindering real-time application performance.
This research proposes a novel control law combined with the PJM algorithm to develop a strategy for full-pose (position and orientation) control of a manipulator’s end-effector. Specifically, a Data-Driven Kinematic Model (DDKM) was developed that achieves rapid convergence of the estimation error, facilitating swift and accurate estimation of the analytical Jacobian matrix. Estimation is carried out using the PJM algorithm as the update policy for the initial Jacobian matrix, requiring minimal knowledge for initialization, specifically the direction of the actuators’ action axes relative to any coordinated frame. Furthermore, several initialization strategies have been proposed in Morales-Díaz et al. [25], ensuring the preservation of important properties of the Jacobian matrix, such as the full-column rank, thereby maintaining invertibility. A key highlight of this work is the ability to estimate an equivalent analytical Jacobian matrix without needing to parameterize the orientation partition using a specific orientation representation, as is typically required in model-dependent control schemes. Instead, by employing pose change rates as feedback for the PJM algorithm, the estimated Jacobian becomes suitable for any orientation representation. This offers increased implementation flexibility as changes in the orientation representation do not necessitate offline modifications to the algorithm. For this research, the simulation is restricted to the use of a similar sensor as the motion capture system, which provides an inertial reference frame. This differs from the case of the handheld camera, where there is no inertial frame, and visual servo control is required. Finally, it is important to emphasize the versatility of the proposed control scheme, proving that it can be implemented in any robot that forms a kinematic chain, including those with arborescent designs, provided the robot’s DOF exceeds the task/constraint DOF. The performance of the DDKM is validated through simulations of both pose regulation and trajectory tracking for various types of manipulator. To test the robustness of the closed-loop configuration, trajectory tracking under disturbances is also performed. A general stability analysis using Lyapunov conditions guarantees the convergence of estimation and control errors, providing guidelines for the design parameters of the DDMC.
The organization of this paper is as follows: Section 2 explains the data-driven kinematic model, Section 3 introduces the novel proposed control law for the robot’s complete pose, Section 4 analyzes the results, and Section 5 covers the conclusions derived from the study.

2. Data-Driven Kinematic Model

The critical pose of the body of a robot, typically the end-effector or its tool, including both its position and orientation, is commonly described as a transformation matrix in 3D space (Lynch and Park [26]),
T ( q i ) = R ( q i ) p ( q i ) 0 1 S E ( 3 ) ,
where T R 4 × 4 is an homogeneous transformation matrix, with p = x , y , z T R 3 , and R S O ( 3 ) is a matrix of the special orthogonal group in three dimensions; thus, R T R = I R 3 × 3 . From (1), for each joint q i in frame i, there is a T i ( q i ) for i = 1 n , where T ( q ) = T 1 ( q 1 ) T 2 ( q 2 ) T n ( q n ) denotes a composite transformation of coordinates, where T ( q ) describes the pose of the end-effector, as a function of its joint coordinates, q R n , where n represents the DOF. It is noteworthy that the orientation representation of T ( q ) is singularity-free, as described by R ( q ) ; however, it is over-parameterized by nine parameters, with six of them being constraints, to represent 3 DOF, which is unsuitable for control purposes. Alternative ways of representing T ( q ) are Euler angles, the axes angle, and the quaternions. From the chosen representation, the pose may take the form χ = p T , θ m i n T T R m , where θ m i n S represents the orientation represented with minimal terms; thus, m = 6 for full-pose control of the end-effector.
The velocity or second-order kinematics describes the end-effector linear and angular velocities, ν = v T , ω T T R 6 , with respect to the robot’s joint velocity, q ˙ R n . However, the twist ν s e ( 3 ) is the Lie algebra of T ( q ) S E ( 3 ) , which is different from χ ˙ = p ˙ T , θ ˙ m i n T T , where p ˙ = v = p ( q ) q q ˙ = J ν q ˙ but θ ˙ m i n ω . Note that R ˙ = R S ( ω ) ; thus, S ( ω ) = R T R ˙ , where R ˙ = R ( q ) q q ˙ , then S ( ω ) = R T R ( q ) q q ˙ . S ( ω ) denotes a skew-symmetric matrix, with S ( ω ) = ω R 3 and ω = ( R T R ( q ) q q ˙ ) , where R ( q ) q R 3 × 3 × n ; thus, R ( q ) q i R 3 × 3 , where R T R ( q ) q i is skew-symmetric; thus, ω = i = 1 n ( R T R ( q ) q i ) q ˙ i , which, similarly to Lynch and Park [26], has the following matrix form:
ω = ( R T R ( q ) q 1 ) ( R T R ( q ) q n ) q ˙ 1 q ˙ n = J ω q ˙ .
The geometric end-effector’s Jacobian is J = J ν , J ω T , such that ν = J q ˙ . Note that the angular velocity ω does not represent the orientation’s change rate in minimal terms, such as θ ˙ m i n , which can be integrated into θ m i n to fully describe the orientation. Consider the following map, ω = J θ θ ˙ m i n , which relates the rotation rates of a given orientation representation and the angular velocity. An inverse map θ ˙ m i n = J θ 1 ω exists if J θ 1 is invertible. Also, let θ ˙ m i n = J o q ˙ , where J o = J θ 1 J ω . Thus, a map χ ˙ q ˙ is carried out as follows:
χ ˙ = I 0 0 J θ 1 J ν J ω q ˙ = J A q ˙ ,
where J A is the analytical Jacobian. The shape of J θ depends on the orientation representation chosen for θ m i n . Let a data-driven kinematic model be expressed as χ * q = J A * R m × n . If χ * , χ ˙ * , and q ˙ are measurable, the only unknown parameter to identify would be J ^ A . To estimate J ^ A , the PJM algorithm introduced in Hou and Jin [27] is applied. This iterative approach updates J ^ A ( k ) based on observations of χ ˙ * ( k ) and an initial guess, J ^ A ( 0 ) , so that J ^ A ( k ) J A * ( k ) as k . It is important to note that a nonsquare J ^ A is estimated, which must remain invertible. The update for J ^ A ( k + 1 ) is given by:
J ^ A ( k + 1 ) = J ^ A ( k ) + η χ ˙ ( k ) * J ^ A ( k ) q ˙ ( k ) q ˙ T ( k ) μ + q ˙ ( k ) 2 .
Since forward kinematics do not need to be explicitly computed, and no prior model information is necessary, the estimation of J ^ A ( k ) is highly computationally efficient (dominant term gives an overall computational complexity of O ( m n ) per iteration). Therefore, for the DDMC approach, the following assumptions need to be satisfied:
Assumption 1.
Lipschitz condition: (4) is a rank-1 update algorithm that improves the estimation of J ^ A ( k ) over iterations by ensuring that χ ˙ * ( k ) J ^ A ( k ) q ˙ ( k ) 0 . Thus, it is assumed that the changes in J ^ A with respect to the input q ˙ are described by constant bounded values that depend on μ and η R + . Namely,   J ^ A ( k + 1 )       L     q ˙ ( k )   , which leads to   χ ˙ * ( k )       L     J ^ A ( k ) q ˙ ( k )   .
Assumption 2.
Assuming that the observed feedback variable for (4) fulfills χ * ( k ) C 1 ( R ) and χ * ( k ) : R R , k R + . Consequently χ ˙ * ( k ) C 1 ( R ) and χ ˙ * ( k ) : R R , k R + . Therefore, it is safe to state that χ ˙ * ( k ) = J * ( k ) q ˙ ( k ) , k > 0 is an actual observation of the real robot behavior, and J * is the “true” Jacobian. Then, for Δ J ( k ) = J ^ ( k + 1 ) J ^ ( k ) , it is assumed that Δ J ( k ) = Δ J ( 0 ) e f ( η , μ ) k , which leads to J ^ ( k ) = J * ( k ) when Δ J ( k ) = 0 , which means that our estimated Jacobian becomes the true Jacobian.

3. Robot Control

Throughout this section is presented a novel closed-loop controller. The control law is designed for regulation and trajectory tracking tasks of the robot end-effector. Once the control law has been established, the stability analysis of the closed-loop controller is demonstrated by the Lyapunov conditions.

3.1. Control Law Design

Assuming that the observed real behavior of the end-effector can be approximated by:
χ ˙ * ( t ) = J A * ( t ) q ˙ ( t ) + ϵ ( t ) q ˙ ( t ) ,
where ϵ ( t ) q ˙ ( t ) are unknown effects caused by the uncertain knowledge of the Jacobian matrix. Let us transform (5) into the discrete domain:
Δ χ * ( k ) δ t = J A * ( k ) Δ q ( k ) δ t + ϵ ( k ) Δ q ( k ) δ t ,
Δ χ * ( k ) = J A * ( k ) Δ q ( k ) + ϵ ( k ) Δ q ( k ) ,
where Δ χ * ( k ) = δ t χ ˙ * ( k ) , Δ q ( k ) = δ t q ˙ ( k ) , and δ t is the sampling time. Also, Δ χ * ( k ) = χ * ( k + 1 ) χ * ( k ) and Δ q ( k ) = q ( k + 1 ) q ( k ) . Considering these changes, (6) becomes:
χ * ( k + 1 ) = χ * ( k ) + J A * ( k ) Δ q ( k ) + ϵ ( k ) Δ q ( k ) .
Since J A * ( k ) is the true Jacobian and ϵ ( k ) is an uncertainty term, J ^ A ( k ) = J A * ( k ) + ϵ ( k ) describes our estimated Jacobian matrix, with a vanishing term ϵ ( k ) according to (4). Therefore, (7) becomes:
χ * ( k + 1 ) = χ * ( k ) + J ^ A ( k ) Δ q ( k ) .
In a general sense, (8) is valid when the orientation representation is minimal and in such a form that its update can be done through discrete summations in a Euler-like integration.
Proposition 1.
If θ S 3 is the orientation represented with quaternions as θ = θ w , θ v T T R 4 , where θ w and θ v = θ x , θ y , θ z T are the scalar and vector terms, respectively, then only the vector part of the quaternion represents the orientation in minimal terms; thus, if the orientation can be updated by a discrete Euler-like integration—just like the position is updated—then the full-pose (8) can be updated as:
p * ( k + 1 ) = p * ( k ) + J ^ ν ( k ) Δ q ( k ) ,
θ m i n * ( k + 1 ) = θ m i n * ( k ) + J ^ o ( k ) Δ q ( k ) .
Proof. 
The discrete quaternion update (10) is derived from the continuous map:
θ ˙ = 1 2 Q ( θ ) ω = 1 2 Q ( θ ) J ^ ω q ˙ ,
where Q ( θ ) ω = θ ω q , and ω q = 0 , ω T T is the angular velocity ω represented as a pure quaternion ω q , and ⊗ denotes a quaternion product. Therefore, Q ( θ ) R 4 × 3 allows one to perform a quaternion product between quaternion θ and 3D velocity vector ω . Let us consider the following quaternion discrete-time update to be an approximation of (11):
θ * ( k + 1 ) θ * ( k ) + 1 2 Q ( θ ) J ^ ω ( k ) Δ q ( k ) ,
The update must be normalized to keep the unit quaternion form θ * ( k + 1 ) θ * ( k + 1 ) . However, if it is defined as J Q = 1 2 Q ( θ ) J ω ( k ) R 4 × n to perform θ ˙ = J Q q ˙ , the analytical Jacobian of (3) would change to J Q A R 7 × n , where m = 6 is the minimal dimension to describe the full pose, which would now become χ Q R 7 . To truly control the pose in minimal terms, the objetive is to control the vector part of the quaternion θ v ; consequently, its update becomes:
θ v * ( k + 1 ) = θ v * ( k ) + 1 2 J θ T J ^ ω ( k ) Δ q ( k ) ,
where J o = 1 2 J θ T J ^ ω ( k ) ; thus, (13) is equivalent to (10). Therefore, the analytical Jacobian for a pose with minimal terms is exactly like (3), and the full pose χ ˙ can be integrated as (8).
The pose errors at step k + 1 become:
e p ( k + 1 ) = p * ( k + 1 ) p d ( k + 1 ) ,
e θ ( k + 1 ) = θ * ( k + 1 ) θ d ( k + 1 ) ,
where θ d = θ w d , θ v d T is the conjugate of θ d . Note that computing the errors in terms of quaternion products requires the complete quaternion θ ; therefore, θ v can be used to reconstruct θ and then normalize it to ensure its unitary property. In this case, the orientation error e θ ( k + 1 ) accounts for the rotation necessary for θ ( k + 1 ) to minimize e θ ( k + 1 ) , which happens when e θ ( k + 1 ) q I , where q I = 1 , 0 , 0 , 0 T is the quaternion identity. By substituting (9) and adding p d ( k ) p d ( k ) = 0 to (14), it is straightforward to obtain the position error update:
e p ( k + 1 ) = e p ( k ) + J ^ ν ( k ) Δ q ( k ) + Δ p d ( k ) ,
where e p ( k ) = p ( k ) p d ( k ) is the position error at iteration k and Δ p d ( k ) = p d ( k ) p d ( k + 1 ) = δ t p ˙ d ( k ) accounts for changes in the desired trajectory. The orientation update may be tricky, as the orientation error requires the full quaternion product (15) and the orientation-state update is done with only the vector part of the quaternion (13). Additionally, it requires multiplying by θ d ( k ) θ d ( k ) = q I to identify e θ ( k ) and Δ θ d ( k ) ; consequently, even combining (15) with (12) is problematic, because it ends up with a product like J Q ( k ) Δ q ( k ) θ d ( k + 1 ) , where J Q ( k ) Δ q ( k ) is not a true unit quaternion.
Proposition 2.
The multiplicative discrete integration of θ ( k + 1 ) can lead us to the multiplicative discrete integration of e θ ( k + 1 ) , in order to obtain a proper orientation error update that mirrors (16) then simplify its multiplicative discrete integration into the following Euler-like summation of the vector parts of the quaternion errors:
e v θ ( k + 1 ) e v θ ( k ) + J ^ o ( k ) Δ q ( k ) + Δ θ v d ( k ) .
Proof. 
The multiplicative discrete update of the orientation can be derived from (11) as follows:
θ ( k + 1 ) = Δ Q ( k ) θ ( k ) ,
where Δ Q becomes
Δ Q ( k ) = exp ω ( k ) δ t 2 ϖ ( k ) = cos ω ( k ) δ t 2 , sin ω ( k ) δ t 2 ϖ ( k ) T ,
where ϖ = ω ω . By substituting (18) into (15), the error update becomes:
e θ ( k + 1 ) = Δ Q ( k ) θ ( k ) θ d ( k + 1 ) ,
then by multiplying it by the identity θ d ( k ) θ d ( k ) = q I , it becomes:
e θ ( k + 1 ) = Δ Q ( k ) θ ( k ) θ d ( k + 1 ) ( θ d ( k ) θ d ( k ) ) .
Regrouping terms using associative properties of the quaternion product leads to:
e θ ( k + 1 ) = ( θ ( k ) θ d ( k ) ) ( Δ Q ( k ) θ d ( k ) θ d ( k + 1 ) ) .
It is now possible to identify e θ ( k ) = θ ( k ) θ d ( k ) and Δ θ d ( k ) = θ d ( k ) θ d ( k + 1 ) . Finally, the error update becomes:
e θ ( k + 1 ) = e θ ( k ) Δ Q ( k ) Δ θ d ( k ) .
If the angular velocities are small, linearization is possible for the sin and cos functions, with a first-order approximation from Taylor series expansions on the exponential map (19) as follows:
Δ Q ( k ) 1 , δ t 1 2 ω ( k ) T .
Therefore, considering that ω = J ^ ω q ˙ and q ˙ = Δ q ( k ) δ t , (23) becomes:
e θ ( k + 1 ) e θ ( k ) 1 , 1 2 J θ T ( k ) J ^ ω ( k ) Δ q ( k ) T Δ θ d ( k ) .
Note that J θ T = I when linearizing (19). The quaternion update e θ ( k + 1 ) is a composite product of quaternions. Also, for the product of two quaternions, θ 1 θ 2 , its vector parts can be simplified as follows: v ( θ 1 θ 2 ) = θ ω 1 θ v 2 + θ ω 2 θ v 1 + θ ω 1 × θ ω 2 θ ω 1 + θ ω 2 ; by following this, e v θ ( k + 1 ) from (25) becomes:
e v θ ( k + 1 ) e v θ ( k ) + 1 2 J θ T ( k ) J ^ ω ( k ) Δ q ( k ) + Δ θ v d ( k ) ,
which, in the continuous time domain, becomes:
Δ e v θ ( k ) δ t 1 2 J θ T ( k ) J ^ ω ( k ) Δ q ( k ) δ t + Δ θ v d ( k ) δ t ,
e ˙ v θ ( k ) 1 2 J θ T ( k ) J ^ ω ( k ) q ˙ ( k ) + θ ˙ v d ( k ) .
In the continuous time domain, the orientation error update becomes the orientation error kinematics, where the angular velocity terms represented in quaternions can be treated as Euclidean vectors and be integrated by a Euler-like summation. Also, note that, in the continuous time domain, J θ does not need to be linearized and can still be integrated like (26). Thus, for a linear discrete-time integration like (26) and assuming a continuous time equivalence to (28), define J θ = I θ w + S ( θ v ) , which lead us to:
e v θ ( k + 1 ) e v θ ( k ) + J ^ o ( k ) Δ q ( k ) + Δ θ v d ( k ) ,
where J ^ o ( k ) = 1 2 J θ T ( k ) J ^ ω ( k ) is the only non-linear term. Therefore, (29) is exactly like (17). This concludes the proof.
Now, consider the following error function, with errors ( k + 1 ) and k:
s ( k + 1 ) = α 1 e ( k + 1 ) + α 2 e ( k ) ,
where α 1 and α 2 are positive constants. By substituting (16) and (17) into (30), we obtain:
s p ( k + 1 ) = α 1 ( e p ( k ) + J ^ ν ( k ) Δ q ( k ) + Δ p d ( k ) ) + α 2 e p ( k ) ,
s o ( k + 1 ) = α 1 ( e v θ ( k ) + J ^ o ( k ) Δ q ( k ) + Δ θ v d ( k ) ) + α 2 e v θ ( k ) .
Note that, since e v θ ( k ) can be integrated like e p ( k ) , we can stack χ = p T , θ v T T , e = e p T , e v θ T T , J ^ A = J ^ ν , J ^ o T , and s = s p T , s o T T . Now, let us use s ( k + 1 ) as a constraint for the following constrained quadratic program:
minimize Δ q k 1 2 ( Δ q ( k ) Δ z ( k ) ) T ( Δ q ( k ) Δ z ( k ) ) s . t . s ( k + 1 ) = 0 ,
which leads to the following objective function:
ξ ( k ) = 1 2 (   Δ q ( k )   2   +     Δ z ( k )   2 ) Δ q T ( k ) Δ z ( k ) λ T ( k + 1 ) s ( k + 1 ) .
Then, solve the following partial derivatives:
ξ ( k ) Δ q ( k ) = Δ q ( k ) + Δ z ( k ) α 1 J ^ A T ( k ) λ ( k ) = 0 ,
ξ ( k ) λ ( k ) = s ( k + 1 ) = α 1 ( e ( k ) + J ^ A ( k ) Δ q ( k ) + Δ χ d ( k ) ) + α 2 e ( k ) = 0 ,
where (35) and (36) build the following system:
I J ^ A T ( k ) J ^ A ( k ) 0 Δ q ( k ) λ ( k ) = Δ z ( k ) a e ( k ) Δ χ d ( k ) ,
where a = α 1 α 1 + α 2 . From (37), solving first for λ ( k ) :
λ ( k ) = ( J ^ A J ^ A T ) 1 ( a e ( k ) + J ^ A ( k ) Δ z ( k ) Δ χ d ( k ) ) .
Then, solve for Δ q ( k ) :
Δ q ( k ) = J ^ A ( k ) ( a e ( k ) + Δ χ d ( k ) ) N ^ ( k ) Δ z ( k ) ,
where J ^ A = J ^ A T ( J ^ A J ^ A T ) 1 is the Moore–Penrose right pseudoinverse of J ^ A and N ^ = I J ^ A J ^ A is a null-space projector that introduce joint velocities Δ z ( k ) R n that have a null effect on χ ˙ ( k ) . Note that (39) effectively minimizes the joint-velocities (33) under the constraint defined by (30). The minimization describes the dynamics of the end-effector task:
Δ e ( k ) = a e ( k ) + Δ χ d ( k ) ,
which is then mapped to the joint-space through J ^ A . As is typically the case with most manipulator’s non-square Jacobians, J ^ A has a solution as long as it is full-column-rank, namely, ρ ( J ^ A ) = m . In order to ensure a solution always exist, this work uses a damped J ^ ϱ A , with a regularization factor ϱ :
J ^ ϱ A = J ^ A T ( J ^ A J ^ A T + I ϱ ) 1 .
Since the Jacobian is estimated, an insufficiently defined direction of the Jacobian that correspond to the coordinates of χ ˙ can be compensated by the action of an adaptable scaling factor for the whole mapping R n R m of (40); this is defined as follows:
c i ( k ) = 1   J ^ A i ( k )   2 + ς i ,
with ς i > 0 . Then, let C k R 6 × 6 be a diagonal matrix with:
C k = diag c x ( k ) c y ( k ) c z ( k ) c θ x ( k ) c θ y ( k ) c θ z ( k ) .
Let us establish a behavior to the task dynamics, which adapts to the changes of the estimated Jacobian while updating. Thus, by using the damped pseudoinverse (41), and (43), (40) in (39), Δ q ( k ) must be redefined as:
Δ q ( k ) = J ^ ϱ A ( k ) C ( k ) u ( k ) N ^ ( k ) Δ z ( k ) ,
where
u ( k ) = α * ( k ) e ( k ) + α ( k ) Δ χ d ( k ) .
with α * ( k ) = α ( k ) α 1 α 1 + α 2 . Also, α ( k ) = C 1 ( k ) if Δ J ^ A ( k ) 0 , otherwise α ( k ) = 1 . Finally, the updated joints’ positions are:
q ( k + 1 ) = q ( k ) + Δ q ( k ) = q ( k ) + δ t q ˙ ( k ) .
While the term N ^ ( k ) Δ z ( k ) does not contribute to the main task (45), it contributes null-space joint-excitation q ˙ n u l l ( k ) = N ^ ( k ) Δ z ( k ) to (4), which keeps the update going in case the range-space joint-velocity q ˙ r a n g e ( k ) = J ^ ϱ A ( k ) u ( k ) is insufficient (e.g., if the robot moves slowly).
The closed-loop configuration is depicted in Figure 1. The PJM algorithm estimates the Jacobian matrix derived from the association between the output signals (end-effector velocities) and input signals (joint’s velocities) since the estimation error ϵ ( k ) 0 . The closed-loop controller starts when the control errors in (14) and (15) are generated between the current end-effector pose and the desired set point. The proposed control law considers the minimization of an objective function, treated as a constrained quasi-Lagrangian function, with the error function s ( k + 1 ) from (30) as a constraint, which contains the control parameters α 1 , α 2 to scale the convergence of the control error e ( k ) . During the learning phase, the control law in (45) is adjusted for each end-effector’s coordinate independently of each other, considering the dynamic system from the equivalent Jacobian matrix with adaptive gain c i ( k ) in (42) scaling the i-th row of the Jacobian matrix.
The numerical method used to solve the equations in the simulations presented in this article is the Euler step integration method, with a step size of 0.01. This method was employed to integrate the first-order kinematic system and to numerically compute the end-effector pose for feedback to the estimator. Using the same integration method, consistency is ensured between system integration and Jacobian updates. To account for potential numerical effects, such as integration errors or step size sensitivity, the selected step size was validated to ensure stability and convergence in the simulations. In real-world applications, the Euler step size should be adjusted based on the precision and noise characteristics of the measurement system to balance computational efficiency and precision.

3.2. Control Stability Analysis

This section demonstrates the Lyapunov stability analysis, involving the system dynamics during the model estimation—learning phase—and the convergence of the error after concluding the learning phase, which shows that a lack of model knowledge, i.e., ϵ ( k ) > 0 , is tolerated for large errors, as the control effort is strongly directed at reducing e ( k ) ; then, as e ( k ) 0 , the influence of ϵ ( k ) becomes more significant relative to e ( k ) and the system’s stability depends on the residual dynamics. The interplay between e ( k ) and ϵ ( k ) and its impact on the stability of the system is addressed in detail in Appendix A, from which the parameters in Table 1 were established.
Theorem 1.
Consider a discrete-time system governed by the error dynamics:
e ( k + 1 ) = e ( k ) + J ^ A ( k ) Δ q ( k ) + Δ χ d ( k + 1 ) ,
where J ^ A ( k ) is the estimated Jacobian from (4), Δ q ( k ) are joint positions increments, and Δ χ d ( k + 1 ) is the desired trajectory increment. A fast and successful estimation of J ^ A ( k ) is mediated by 0 < η 2 ( μ + q ˙ ( k ) 2 ) q ˙ ( k ) 2 and μ > ( η + 2 ) q ˙ ( k ) 2 2 , both of which ensure ϵ ( k ) 0 . See Morales-Díaz et al. [25], Gómez et al. [28]. Also, consider Assumptions (1) and (2) and the following Lyapunov function candidate:
V c ( k + 1 ) = 1 2 e T ( k + 1 ) e ( k + 1 ) ,
and its discrete change is given by:
Δ V c ( k + 1 ) = Δ e ( k + 1 ) T e ( k ) + 1 2 Δ e ( k + 1 ) .
The Lyapunov function satisfies Δ V c ( k + 1 ) 0 for all k. Furthermore:
1.
If ϵ ( k ) 0 (during learning), Δ V c ( k + 1 ) < 0 as long as:
a   e ( k )   λ min ( M k C k ) > Γ k a λ max ( M k C k ) + 1 + 1 2 Γ k 2 e ( k ) ,
and e ( k ) > e ( k ) min , where M k = ( J A * ( k ) + ϵ ( k ) ) J ^ A ( k ) encodes uncertainty during the learning process and Γ k = λ max ( G ) Δ χ d ( k + 1 ) + Z k is a disturbance term, with G = M k C k I and Z k = ( J A * ( k ) + ϵ ( k ) ) N ^ ( k ) Δ z ( k ) .
2.
After learning ( ϵ ( k ) 0 ):
Δ V c ( k + 1 ) = a e ( k ) T e ( k ) + 1 2 a 2 e ( k ) T e ( k ) ,
which is strictly negative for 0 < a < 1 .
Thus, the system is globally asymptotically stable, and e ( k ) 0 as k .
Proof: 
The Lyapunov function V c ( k ) = 1 2 e ( k ) T e ( k ) is positive definite. Its discrete-time change is:
Δ V c ( k + 1 ) = a e ( k ) T M k C k e ( k ) + a 2 2 e ( k ) T M k C k M k C k e ( k ) + Ω k T e ( k ) + 1 2 Ω k 2 .
During learning, ( ϵ ( k ) 0 ), Ω k 0 due to the uncertainty in the Jacobian estimation. The dominant negative term a e ( k ) T M k C k e ( k ) ensures Δ V c ( k + 1 ) 0 as long as:
a e ( k ) 2 λ min ( M k C k ) > Γ k a M k C k e ( k ) + e ( k ) + 1 2 Γ k 2 ,
where, with bounded Γ k , e ( k ) , and proper tuning of a, C k , and regularization with ς ensure stability. After learning ( ϵ ( k ) 0 ), M k I and Ω k 0 , the Lyapunov’s discrete change simplifies to:
Δ V c ( k + 1 ) = a e ( k ) T e ( k ) + 1 2 a 2 e ( k ) T e ( k ) ,
which is strictly negative for 0 < a < 1 , ensuring asymptotic convergence e ( k ) 0 . Therefore V c ( k ) is positive definite, radially unbounded, and decreases monotonically ( Δ V c ( k + 1 ) 0 ). Hence, the system is globally stable and the error converges asymptotically to zero. Refer to Appendix A for in-depth details and analysis.

4. Results: Ddkm Control Performance Evaluation

This section is dedicated to evaluating the proposed DDKM scheme. In Section 4.2, a simulation is shown to validate the pose regulation data-based controller of Section 3.1 in (44) and (40). In Section 4.1, a simulation is presented to corroborate the trajectory tracking data-based controller. Finally, in Section 4.3, a simulation is presented to verify the pose regulation with perturbation data-based control law. To evaluate the controller performance, it is used in a CoppeliaSim Edu V4.6.0-Rev16 robotics simulator environment, developed by Coppelia Robotics AG, located in Zurich, Switzerland. Data-based control computations were implemented using Python 3.11.5 on an Asus laptop computer with 6 GB of RAM, Intel® Core ™i7-8550U CPU @ 1.8 GHz × 8, and Ubuntu 20.04.3 LTS 64-bit operating system. For communication between robot and computer, the zmq remote API functions of CoppeliaSim were used. For the first simulation, a 6-DOF inertial robot manipulator from the brand Universal Robots, model UR5, is used. The controller is then tested on a redundant omnidirectional mobile manipulator with 8 DOF from the Kuka-Youbot (KY) brand. For the final test, the Kuka-Youbot omnidirectional mobile base with dual arm (KYD) is used, providing 13 DOF to control the full poses of two end-effectors.
Different desired coordinates for the end-effector’s pose of the simulated robots were assigned. The aim is to challenge the performance of the control scheme, demonstrating how the robot can maneuver its joints to reach a diverse set of coordinate combinations for the end-effector. Throughout the simulation run-time, the desired pose coordinates were interactively varied by moving reference objects. The simulation setup for each robot is shown in Figure 2.
To further evaluate the robustness of the proposed control scheme, initial conditions were carefully chosen to include challenging scenarios, such as starting from singular positions. These cases serve as critical tests for the control system. For other initial conditions, whether singular or non-singular, the control system demonstrates successful convergence.
It is important to note that the robot’s initial orientation relative to the desired pose must be aligned with the z axis, within a range of 0 to 180 degrees. This restriction is due to the inherent sign ambiguity in quaternion representations. Future work could focus on addressing this limitation to enhance the generality of the system.
During simulation run-time, the end-effector of various manipulators was controlled by interactively adjusting the desired pose coordinates. The multimedia material at the end of each experiment displays the maneuvers performed in these simulations.

4.1. Inertial Manipulator

When the manipulator is fixed on its base, it has limited mobility and reach, resulting in a reduced set of compatible position and orientation coordinates that can be assigned. Figure 3 presents the pose commanded for the UR5 robot during 250 s of the simulation. Subfigure (a) displays the position coordinates (x, y, z) compared to the desired values ( x d , y d , z d ) over time. It is evident that the position commands were dynamically adjusted, reflecting significant changes in the z-axis around the 75 s and 150 s marks. The x and y coordinates follow a smoother trajectory, although slight deviations from the desired positions can be observed, particularly in the y-axis after 200 s. Subfigure (b) depicts the commanded orientation of the end-effector, represented as a quaternion ( θ ω , θ x , θ y , θ z ). The orientation follows a more complex behavior with abrupt changes, particularly in θ x and θ z , around the same time intervals as the positional changes. These variations indicate rotational adjustments made to maintain the desired end-effector pose, with the quaternion components stabilizing towards the end of the simulation.
Figure 4 illustrates the absolute errors of position (14) and orientation (15) of the UR5 end-effector throughout the simulation. Subfigure (a) shows the position errors for the coordinates x, y, and z. Initially, there are noticeable fluctuations, particularly in the z-axis, but these errors quickly converge to values close to zero, indicating an effective stabilization of the system. Small deviations can still be observed during moments where the commanded pose changes, particularly around the 50, 100, and 150 s marks, reflecting the system’s response to new target positions. Subfigure (b) presents the orientation errors, represented as quaternion components ( e θ ω , e θ x , e θ y , e θ z ). Similarly to position errors, orientation errors stabilize over time. In particular, the quaternion component e θ ω , associated with the real part, remains close to one, indicating a stable rotation. The imaginary components ( e θ x , e θ y , e θ z ) exhibit transient variations but converge to near-zero values, ensuring that the end-effector maintains its desired orientation as commanded. The system remains stable until significant changes in the commanded orientation are introduced.
Figure 5 illustrates the end-effector velocities of the UR5 robot, derived from the time derivative of the pose coordinates captured in the simulation. Subfigure (a) displays the linear velocities along the x-, y-, and z-axes. Initially, there are notable fluctuations, particularly in the x direction, where the velocity reaches peaks of approximately ±1.5 m/s during the transitions. These fluctuations indicate the system’s response to the commanded poses, which subsequently stabilize as the motion becomes smoother over time. Subfigure (b) presents the angular velocities represented by the quaternion rates ( θ ˙ x , θ ˙ y , θ ˙ z ). Similarly to linear velocities, the angular rates exhibit initial peaks, particularly around the transition phases. As the simulation progresses, these rates stabilize and converge toward zero, indicating that the end-effector effectively maintains its orientation without excessive rotation. This behavior underscores the efficiency of the control system in achieving stable movement in both linear and angular dimensions.
Figure 6a illustrates the evolution of the adaptive gains for each element of the end-effector’s pose. These gains start from an initial value near 0.5 and evolve dynamically as the controller adapts to changes in the task, reaching values between 0.4 and 2. In particular, the gains associated with the orientation components ( c θ x , c θ y , and c θ z ) show significant variations when the orientation commands change, indicating the controller’s response to orientation uncertainties. The position gains ( c x , c y , and c z ) exhibit more gradual variations, reflecting smaller positional uncertainties throughout the trajectory. Figure 6b shows the joint velocities of the UR5 robot as the control output. The velocities remain continuous and stable, with values ranging between −2 and 2 rad/s for all joints. Peaks in velocity can be observed at the beginning of the task and during key moments when the desired pose coordinates are altered. These peaks are consistent across all joints, indicating synchronized movement. Despite these peaks, the joint velocities remain well within the operational safety limits, ensuring smooth robot motion without abrupt changes. These peaks can be handled with smooth transition functions, which take values from 0 to 1; these would be applied at the moments when the target pose changes are generated, as done in Toro-Arcila et al. [29]. The simulation for this experiment is presented in a video available at https://drive.google.com/file/d/12vOkFRL-_AVmSgVu3xISPHUkiIzejoGR/view?usp=sharing (accessed on 30 October 2024).

4.2. Omnidirectional Mobile Manipulator

When the manipulator has an omnidirectional mobile base, it gains an additional 3 DOF because of its holonomic constraint, which makes it a redundant manipulator. Also, its mobility and reach are increased; therefore, it has a wider set of feasible coordinates for the end-effector to reach.
Figure 7 shows the evolution in time of the commanded position and orientation of the end-effector during the task (200 s of simulation). In Figure 7a, the position coordinates ( x d , y d , z d ) are interactively modified at specific moments during the simulation. The component x d shows a significant change at around 150 s, while z d decreases slightly and y d remains relatively stable throughout the simulation, with minimal fluctuations. In Figure 7b, the commanded orientation, expressed as quaternion components ( θ ω , θ x , θ y , θ z ), also undergoes dynamic changes. Significant variations occur in θ x and θ y , especially during the first 50 s, reflecting adjustments in the desired orientation. Approaching the end of the simulation, the quaternion values stabilize, indicating a steady desired orientation after the commanded modifications.
Figure 8 illustrates the evolution in pose errors in time for the KY end-effector. In subfigure (a), the position errors ( e x , e y , e z ) initially exhibit significant discrepancies, especially in the z-axis, but rapidly converge to near-zero values as the system adapts. This rapid reduction in error demonstrates the effectiveness of the control algorithm in aligning the actual position with the commanded values. However, small oscillations can be observed around t = 100 s, particularly on the z-axis, which correspond to minor fluctuations likely caused by dynamic disturbances or changes in the commanded pose. Subfigure (b) depicts the orientation errors represented by the imaginary parts of the unit quaternion components ( e θ ω , e θ x , e θ y , e θ z ). The quaternion errors follow a similar trend, with a rapid reduction in discrepancies over time. Although the errors converge close to zero, there are slight variations, especially in e θ x and e θ y , reflecting the challenge of maintaining precise orientation under dynamic conditions. These variations remain within acceptable limits, indicating the overall robustness of the system in handling orientation changes. Both subfigures confirm that the control approach ensures the convergence of the pose errors over time, with the position and orientation errors stabilizing near zero until the introduction of new commanded pose coordinates, where a temporary increase in error occurs before reconvergence.
Figure 9 presents the linear and angular velocity profiles of the end-effector. Subfigure (a) shows the linear velocities x ˙ , y ˙ , and z ˙ in the Cartesian coordinate system, while Subfigure (b) displays the angular velocity rates θ ˙ x , θ ˙ y , and θ ˙ z , expressed in terms of quaternions. These velocities were obtained by numerically differentiating the pose data captured from the simulation environment. The linear velocity plot shows significant variations in the early stages, followed by stabilization, which aligns with the expected transient behavior of the system. Similarly, the quaternion rates in Subfigure (b) illustrate the dynamic changes in orientation over time. Both sets of velocity serve as feedback for the adaptive control algorithm, enabling the system to correct deviations in real time by adjusting the control inputs. This feedback loop ensures precise motion control of the end-effector throughout the simulation period.
Figure 10a depicts the evolution of the adaptive gains for the controller, applied to the individual elements of the complete pose. The adaptive gains, denoted as c x , c y , c z for the Cartesian coordinates and c θ x , c θ y , c θ z for the quaternions, vary within the range of 0.8 to 2.0 , particularly under conditions of maximum uncertainty. These gains adjust dynamically in response to changes in the system’s state, converging to stable values when the desired pose coordinates remain constant. The plot illustrates how the controller adapts to ensure optimal performance over time. Figure 10b illustrates the joint angular velocities q ˙ 1 to q ˙ 8 , representing the control output for each of the manipulator’s joints. The velocities remain within feasible and safe operational limits and exhibit smooth transitions without abrupt discontinuities. These control signals ensure that the manipulator operates within a stable range, maintaining the safety and efficiency of the system. The combined data from both plots demonstrate the robustness of the adaptive control strategy in handling both position and orientation tracking. The simulation for this experiment is presented in a video available at https://drive.google.com/file/d/1bhFaYJZR0S9HDmpxmqKqhFtpigMTwi0H/view?usp=sharing (accessed on 30 October 2024).

4.3. Dual-Arm Omnidirectional Mobile Manipulator

When the omnidirectional mobile base is equipped with a dual system of manipulators, there are two end-effectors to control and additional degrees of freedom (DOF), totaling 13 DOF. Of these, three are shared (omnidirectional mobile base), while each manipulator contributes five exclusive DOF for pose control. In this case, it becomes important to assign feasible pose coordinates to both end-effectors, which must remain near their mobile base.
Figure 11 and Figure 12 present the commanded pose coordinates for the left and right end-effectors, which were modified during simulation time (110 s). Figure 11a shows the commanded position trajectories for the left tip of the manipulator in Cartesian space, with position coordinates x l , y l , z l , as well as their desired references x l d , y l d , z l d . The plot highlights the convergence of the actual position coordinates to the desired values over time. It is evident that the system rapidly stabilizes the x l coordinate, while minor oscillations can be observed in the y l and z l coordinates before reaching steady state. These variations reflect the system’s response to the initial command and subsequent adjustments, achieving precise position control under dynamic conditions. Figure 11b presents the commanded orientation based on quaternions for the left tip, including the actual components of the quaternions θ l ω , θ l x , θ l y , and θ l z , along with their desired references θ l ω d , θ l x d , θ l y d , and θ l z d . The system demonstrates smooth transitions between the desired orientations, with notable jumps in the quaternion components corresponding to large rotational adjustments. The data in both plots demonstrate the manipulator’s ability to achieve accurate position and orientation tracking, even when subjected to complex commands. This precise control is crucial for tasks that require high levels of dexterity and adaptability in robotic manipulation.
The results presented in Figure 12 illustrate the evolution of the commanded pose for the right end-effector, including both position and orientation, over a simulation period of 110 s. The figure is divided into two subplots: position (Figure 12a) and orientation (Figure 12b). In position subplot (a), the actual and commanded Cartesian coordinates, x, y, and z, of the right end-effector are displayed over time. Solid red, green, and blue lines represent the actual values of x r , y r , and z r , while the corresponding dashed lines represent the desired trajectories, x r d , y r d , and z r d . The coordinate x (red line) shows a rapid rise at the beginning of the simulation, reaching the desired value of approximately 2 m within the first 10 s. This initial transient phase indicates that the controller successfully manages the acceleration and deceleration necessary to achieve the target. After reaching the desired value, the position x stabilizes with minimal overshoot and remains steady until the end of the simulation, showing no significant deviation from the reference trajectory. The coordinates y (green line) and z (blue line) exhibit a more complex behavior. The coordinate y shows an initial negative deviation before converging to zero. Around the 80 s mark, a minor disturbance occurs, likely due to changes in the commanded orientation, which is quickly corrected, indicating robust control. The coordinate z remains largely stable, fluctuating around the desired value with only small deviations, particularly between 40 and 60 s. These deviations are due to changes in the commanded orientation but are effectively controlled to maintain the end-effector within the desired values. In orientation subplot (b), the four quaternion components ( θ r ω , θ r x , θ r y , and θ r z ) are shown. Initially, all quaternion components deviate significantly from their desired values, reflecting the misalignment of the end-effector at the start of the simulation. However, as time passes, the controller adjusts the orientation smoothly. Between 10 and 30 s, there is a noticeable shift in all components, with the system gradually aligning with the target orientation. The most prominent change occurs in θ r y (yellow line) and θ r z (cyan line), which undergo significant transitions before settling in their desired values. By the 40 s mark, the quaternion components stabilize, indicating that the end-effector has successfully aligned with the commanded orientation. The transient behavior observed in the position and orientation plots suggests that the control strategy employed in the KYD robot is effective in handling both position and orientation. The system demonstrates a quick response to initial errors, as evidenced by the rapid convergence to the desired pose, while also showing resilience to disturbances encountered mid-simulation, particularly in the y and z coordinates. Overall, the results suggest that the controller is capable of achieving precise and stable control over the end-effector’s position and orientation, making it a robust candidate for tasks requiring high precision in dynamic environments.
Figure 13 and Figure 14 display the pose errors for the left and right end-effectors, respectively, including position and orientation errors. The graphs reveal a consistent trend of error convergence towards zero over time, signifying effective control despite initial deviations. In Figure 13a, the position errors for the left end-effector, represented by e x l (red), e y l (green), and e z l (blue), show a rapid reduction to near zero values within the first 10 s of the simulation. The largest initial error is observed in e x l , where the deviation reaches nearly −2 m at the beginning but quickly stabilizes. Similarly, e y l and e z l also converge, demonstrating smooth error correction over time. Small oscillations are noticeable around 50 s, particularly in e y l and e z l , which are a consequence of changes in the desired position. However, these deviations are promptly corrected, leading to stable performance in the later stages of the simulation. The quaternion errors that represent the orientation of the left end-effector are shown in Figure 13b. Initially, significant deviations are observed in all quaternion components ( e θ l ω , e θ l x , e θ l y , e θ l z ). These errors gradually decrease, with the largest deviations in e θ l ω (black) and e θ l y (yellow). Between 20 and 60 s, fluctuations in e θ l x and e θ l y indicate small orientation adjustments. By the end of the simulation, the quaternion errors have reduced substantially, reflecting that the left end-effector aligns closely with the desired orientation.
In Figure 14a, the position errors for the right end-effector follow a similar trend. The e x r (red) error starts at approximately −2 m, the largest deviation among the three axes, but rapidly decreases within 10 s. The errors e y r (green) and e z r (blue) also exhibit a fast convergence, with only minor deviations throughout the simulation. The convergence of position errors demonstrates the effectiveness of the controller in maintaining the desired end-effector positions. The orientation errors for the right end-effector, depicted in Figure 14b, show a pattern similar to that of the left end-effector. The quaternion errors initially present significant deviations, especially in e r θ ω (black) and e r θ y (yellow). Although small fluctuations are observed between 40 and 80 s, particularly in e r θ x and e r θ y , the system exhibits stable behavior in the final phase of the simulation, with all quaternion errors converging toward zero.
In both cases, the results clearly demonstrate that the controller effectively reduces both position and orientation errors over time, ensuring accurate tracking of the end-effector poses. The system shows rapid error convergence during the first 10 s and remains robust against disturbances (changes in the desired values of the position and orientation of both end-effectors), particularly in the middle stages of the simulation. This performance highlights the controller’s ability to adapt to changing reference coordinates while maintaining high precision and stability.
In Figure 15 and Figure 16, the velocity profiles of the left and right end-effectors are presented, detailing their linear and angular components over time. These velocity profiles were obtained from the simulation environment and serve as critical feedback for the adaptation algorithm used in the control strategy. In Figure 15, the linear velocities of the left end-effector along the axes x, y, and z are shown in subplot (a). It can be observed that the velocities stabilize after an initial transient phase, converging close to zero after approximately 10 s. The x-axis velocity ( x ˙ l ) exhibits minor oscillations initially but rapidly attenuates to near zero values. Similarly, the y-axis velocity ( y ˙ l ) stabilizes, while the z-axis velocity ( z ˙ l ) maintains minimal fluctuations throughout the simulation. Subplot (b) of Figure 15 depicts the angular velocities of the left end-effector, represented as quaternion rates θ ˙ l x , θ ˙ l y and θ ˙ l z . These quaternion rates capture the rotational dynamics, showing an initial adjustment period where the angular velocities exhibit greater variance, particularly in the θ ˙ l z component. After approximately 10 s, the angular velocities stabilize, with θ ˙ l z showing minor oscillations, while θ ˙ l x and θ ˙ l y converge towards zero.
Similarly, Figure 16 presents the linear and angular velocities of the right end-effector. Subplot (a) shows the linear velocities along the x-, y-, and z-axes. The velocity of the x-axis ( x ˙ r ) of the right end-effector follows a trend similar to that of the left end-effector, showing initial oscillations before stabilizing to near zero after 10 s. The velocity of the y-axis ( y ˙ r ) also stabilizes, although with slightly higher transient peaks compared to the left end-effector. The velocity of the z-axis ( z ˙ r ) remains stable, with minor deviations throughout the simulation. Subplot (b) in Figure 16 displays the quaternion rates θ ˙ r x , θ ˙ r y , and θ ˙ r z , corresponding to the angular velocities of the right end-effector. Similarly to the left end-effector, the quaternion rates exhibit an initial adjustment phase, particularly in θ ˙ r x and θ ˙ r z . The angular velocities then stabilize, with all three components converging to near-zero values after approximately 10 s.
Overall, the velocity profiles for both the left and right end-effectors indicate effective stabilization of both linear and angular velocities over time, highlighting the robustness of the control algorithm in adapting to dynamic conditions. The transient oscillations observed in the first 10 s are typical of systems with high sensitivity to initial conditions, but the convergence to near-zero velocities demonstrates that the system successfully achieves stable motion control of both end-effectors.
Figure 17 illustrates the adaptive control gains and joint velocities for both left and right end-effectors, providing insight into the controller’s performance under varying values of the desired pose. In Figure 17a,b, the adaptive gains for the left and right pose controllers are presented, respectively. The adaptive gains for the left end-effector, shown in Figure 17a, vary within a range of 0.6 to 2, with noticeable adjustments occurring principally when the system experiences greater changes in the desired values of the pose. The gain values ( C l x , C l y , C l z , C l θ x , C l θ y and C l θ z ) initially fluctuate before stabilizing once the system reaches a steady state, and the desired pose coordinates remain unchanged. Similarly, in Figure 17b, the adaptive gains for the right end-effector ( C r x , C r y , C r z , C r θ x , C r θ y and C r θ z ) follow a similar pattern, ranging from 0.5 to 2. These gains exhibit an increase in response to changes in the desired values of the pose and then stabilize as the system converges. Figure 17c depicts the joint angular velocities, representing the control signals generated by the system to drive the manipulator. All these angular velocities ( q ˙ b 1 , q ˙ b 2 , q ˙ b 3 , q ˙ l 1 , …, q ˙ l 5 , q ˙ r 1 , …, q ˙ r 5 ) are continuous and remain within feasible operational ranges, ensuring safe and stable motion control. The velocities show small oscillations in the transient phase, followed by stabilization once the adaptive control mechanism compensates for the changes in the desired values of the pose. The absence of abrupt changes in joint velocities further demonstrates the robustness of the control strategy in maintaining smooth and controlled motion under variable conditions. The simulation for this experiment is presented in a video available at https://drive.google.com/file/d/1Yhp-rIkspLm5suoJIVKH6w81mnRh2ACN/view?usp=sharing (accessed on 30 October 2024).

5. Data-Driven Control Versus Model-Based Control

The performance of the data-drive control scheme is compared against the classical model-based control of the end-effector’s pose of the omnidirectional mobile manipulator Kuka-Youbot. Using either estimated J ^ A or theoretical J A Jacobians has a strong impact on the convergence of e p , e θ and the overall behavior of the end-effector’s dynamics—even affecting the joint-space due to redundancy. The difference between both approaches can be summarized by considering the dependencies on (44) for a data-driven Δ q d d ( k ) and a control based on the modeled inverse-kinematics Δ q m b ( k ) :
Δ q d d ( k ) = J ^ ϱ A ( k , ϵ ( k ) · q ˙ ( k ) ) C k u ( k ) N ^ ( k , ϵ ( k ) · q ˙ ( k ) ) Δ z ( k ) , Δ q m b ( k ) = J ϱ A ( q ( k ) ) u ( k ) N ( q ( k ) ) Δ z ( k ) .
Now, considering a task for the end-effector (5) together with (55), the closed loop system is:
Δ χ d d * ( k ) = M k C k u d d ( k ) Z k ,
Δ χ m b * ( k ) = M ¯ k u m b ( k ) Z ¯ k ,
where M ¯ k = ( J A * ( k ) + ϵ ( k ) ) J A ( k ) I and Z ¯ k = ( J A * ( k ) + ϵ ( k ) ) N k Δ z k 0 if J A ( q ( k ) ) = J A * ( k ) , which requires χ ( q ( k ) ) = χ * ( k ) , such that χ ( q ) q = χ * q . Thus, if χ * is a measured signal of the robot’s pose, it may differ from theoretical χ ( q ( k ) ) if the reference point for sensor measurements is slightly misaligned, namely, χ * ( k ) = χ ( q ( k ) ) + I δ χ ( q ( k ) ) . This leads to χ * q = χ ( q ) q + I δ χ ( q ( k ) ) q , leading to J A ( q ( k ) ) + ϵ m b ( q ( k ) ) = J ˜ A ( q ( k ) ) , where J ˜ A ( q ( k ) ) is a misaligned theoretical Jacobian, where, in contrast to J ^ A ( k ) J A * ( k ) , if J ˜ A ( q ( k ) ) J A * ( k ) it does not adapt. The adaptability of J ^ A ( k ) towards the real observed behavior of the robot, J A * ( k ) , is the strongest advantage of this approach. Note from (55) that, after the learning phase, J ^ ϱ A ( k , ϵ ( k ) · q ˙ ( k ) ) J ^ ϱ A ( k , 0 · q ˙ ) , namely, the Jacobian becomes invariant to q ˙ , while being able to minimize e ( k )
For a comparative analysis, we designed the following trajectory tracking task: χ d ( t ) = [ p d ( t ) , θ d ( t ) ] T , with p d = [ x d ( 0 ) + sin ( t ) , y d ( 0 ) + cos ( t ) , 0.6 ] T , where ( x d ( 0 ) , y d ( 0 ) ) is the center of the circle, θ d = [ 1 , 0 , 0 , 0 ] T . Also, χ ˙ ( t ) = [ p ˙ d ( t ) , θ ˙ d ( t ) ] T , with p ˙ d ( t ) = [ cos ( t ) , sin ( t ) , 0 ] T , θ ˙ d ( t ) = [ 0 , 0 , 0 , 0 ] T . The robot’s initial pose, χ ( 0 ) , in the virtual environment is p ( 0 ) = [ 1.8 , 0 , 0.745 ] T , θ ( 0 ) = [ 1 , 0 , 0 , 0 ] T . The errors are computed with (14) and (15) and the control law is (45). This task simulates a scenario where the robot polishes the surface of a disk from below; see Figure 18. This demands high mobility and precision; the desired end-effector orientation, θ d , is problematic due to the robot structure and redundancy.
The simulation current time is discretized as t = ( 1 k ) δ t , where k = 1 , 2 , , κ , and it lasts a total time of τ = κ δ t . Also, p d ( τ i ) generates a complete circle in τ i = κ i δ t , with κ Z and κ i 2 π δ t . Then, for δ t = 0.01 , κ i = 628 for each cycle i. The simulation lasts 16 cycles for a total duration of κ = 16 κ i = 10048 and τ 100 s . Since the trajectory tasks are repetitive cycles, it is possible to partition the data into sets that correspond to individual cycles, such as χ ¯ 1 = [ χ ( t o ) , χ ( t 1 ) , , χ ( τ 1 ) ] , thus defining χ ¯ 1 , χ ¯ 2 , , χ ¯ 16 , and similarly χ ¯ d 1 , χ ¯ d 2 , , χ ¯ d 16 , such that they can be used to compute performance indexes, such as the Root Mean Square Error: R M S E i ( χ ¯ i , χ ¯ d ) 2 = 1 κ i k = 1 κ i ( χ k χ d ) , the integral of absolute error: I A E ( χ ¯ i , χ ¯ d ) δ t k = 1 κ i χ k χ d , the integral of the squared error I S E ( χ ¯ i , χ ¯ d ) δ t k = 1 κ i χ k χ d 2 , and the integral of Time-weighted Absolute Error: I T A E ( χ ¯ i , χ ¯ d ) δ t k = 1 κ i k δ t χ k χ d . The performance indexes compare the consistency of task resolution for a repetitive task over a relatively long time. The comparison of the data-driven control against the inverse-kinematic control are seen in Figure 19, where the data-driven control scheme consistently outperforms the control based on inverse-kinematics. The nature of this task (particularly the desired orientation) involves a joint configuration that is close to singularity for this robot, which is reflected after the 4th cycle of the task, where the inverse-kinematics control scheme fails to maintain the desired orientation, and the orientation of the end-effector goes off-course, highlighting the susceptibility of the theoretical Jacobian to singular configurations. Nonetheless, during the 2nd and 3rd cycles the inverse-kinematics control scheme fulfilled the task with a decent level of consistency across cycles; however, the data-driven approach performed better, showing a consistent minimization of the errors across every cycle. The simulation for this experiment is presented in a video available at https://drive.google.com/file/d/1xcwY69OdfPCK9jXHNnmtOqmkDJ3sQKgi/view?usp=drive_link (accessed on 30 October 2024).

6. Conclusions

This work proposes a data-driven control strategy for controlling the full pose (position and orientation) of manipulators, which is feasible as long as the dimension of the task coordinates does not exceed the dimension of the joint space. Additionally, the pose and target coordinates must be continuously measured and sufficiently smooth to allow differentiation. An estimation algorithm is used, which relies on an update law and an initial guess of the Jacobian matrix. This approach facilitates the rapid computation of the estimated Jacobian matrix, leading to fast convergence of the estimated signal (end-effector’s velocity) in respect to the measured one. Rapid convergence of the estimation signals ensures the convergence of the controller error terms, as demonstrated in the stability analysis. The stability of the controlled system is validated through a Lyapunov analysis, considering the pose and estimation error terms.
A series of simulations were conducted to test the control performance when applied to different robots, demonstrating the implementation flexibility of this approach, as a no kinematic model needs to be computed, and the geometry and physical parameters of the robot are not required. This facilitated implementation with a variety of robots, including those with an arborescent kinematic chain.
Additionally, the simulations demonstrate how the controller can track desired pose coordinates, which were interactively modified during simulation time. These variations can be considered disturbances in the tracking process. Despite this, the controller successfully stabilized the errors to near-zero values. In the first simulation, an UR5 inertial manipulator with 6 DOF, fully actuated for full-pose control, was used, resulting in a square Jacobian matrix. Next, a Kuka-Youbot with 8 DOF was employed, which is redundant, leading to a non-square Jacobian. Finally, the mobile dual-arm Kuka-Youbot with 13 DOF was used, which is redundant for the full-pose control of both end-effectors. Since this is an arborescent kinematic chain, a non-square block Jacobian matrix was obtained. In all cases, the Jacobian matrices were successfully estimated, achieving the objectives of the proposed control strategy. Notably, this approach is suitable for any kind of angular representation, such as quaternion, Euler angles, and axis angle. In addition, this strategy is applicable to any type of manipulator (inertial or mobile), as long as the conditions mentioned above are satisfied.
Considering the dependency on the quality measurement of the end-effector pose signal, this limits the data-driven approach; however, as future work, a solution method will be applied to measure the end-effector’s pose coordinates using a motion capture systems or on-board devices on the robots to feed the PJM algorithm. These methods would be necessary to implement this control approach in experiments with real robots. In addition, it is planned to include an external force control loop to include other industrial applications.

Author Contributions

Conceptualization, J.G.-C. and J.O.-F.; methodology, J.G.-C., C.A.T.-A. and J.O.-F.; software, J.F.M.-V. and O.G.-C.; validation, N.A.R.-R. and D.E.O.-R.; formal analysis, C.A.T.-A., J.O.-F. and D.E.O.-R.; investigation, J.O.-F. and D.E.O.-R.; resources, N.A.R.-R.; data curation, J.O.-F. and O.G.-C.; writing—original draft preparation, J.G.-C., C.A.T.-A. and D.E.O.-R.; writing—review and editing, J.O.-F. and D.E.O.-R.; visualization, J.G.-C., C.A.T.-A. and J.O.-F.; supervision, J.G.-C. and J.F.M.-V.; project administration, J.G.-C. and N.A.R.-R.; funding acquisition, J.G.-C. and N.A.R.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in the study are included in the article; further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following acronyms and symbols are used in this manuscript:
DDKMData-Driven Kinematic Model
DDMCData-Driven Modeling and Control
DOFDegrees Of Freedom
IFRInternational Federation of Robotics
KYKuka-Youbot
KYDKuka-Youbot Dual
MIMOMultiple-Input Multiple-Output
NNNeural Network
PJMPseudo Jacobian Matrix
RBFRadial Basis Function
ROSThe Robot Operating System
UUBUniformly Ultimately Bounded
THomogeneous transformation matrix in R 4 × 4
pPosition vector in R 3
RRotation matrix in R 3 × 3
q i Joint in reference frame i
χ Pose vector
θ m i n Orientation represented with minimal terms
ν Velocity vector in R 6
vLinear velocity vector in R 3
ω Angular velocity vector in R 3
q ˙ Joints velocity vector
J           Geometric end-effector’s Jacobian
J A Analytical Jacobian
η Adaptation weight parameter
μ Adaptation regularization parameter
ϵ Estimation error
δ t Sampling time
θ Orientation represented with quaternions
θ w Quaternion scalar component
θ v Quaternion vectorial component
e p Position error
e θ Orientation error
α 1 Control gain
α 2 Control gain
α Learning-aware compensation term

Appendix A. Proof of Theorem 1

The discrete-time Lyapunov function for the controller stability is:
V c ( k + 1 ) = 1 2 e T ( k + 1 ) e ( k + 1 ) ,
and the change in the discrete-time Lyapunov function for the controller is defined as:
Δ V c ( k + 1 ) = V c ( k + 1 ) V c ( k ) , = 1 2 e ( k + 1 ) T e ( k + 1 ) 1 2 e T ( k ) e ( k ) ,
where (A2) is equivalent to:
Δ V c ( k + 1 ) = ( e ( k + 1 ) T e ( k ) T ) e ( k ) + 1 2 ( e ( k + 1 ) e ( k ) ) .
By setting Δ e ( k + 1 ) = e ( k + 1 ) e ( k ) , the Lyapunov function in terms of the error difference is:
Δ V c ( k + 1 ) = Δ e ( k + 1 ) T e ( k ) + 1 2 Δ e ( k + 1 ) .
By stacking (16) and (17), the pose error dynamics becomes:
e ( k + 1 ) = e ( k ) + J ^ A ( k ) Δ q ( k ) + Δ χ d ( k ) .
Expand (A5); use e ( k ) and (44) to define Δ e ( k + 1 ) :
Δ e ( k + 1 ) = J A * ( k ) Δ q ( k ) + ϵ ( k ) Δ q ( k ) + Δ χ d ( k + 1 ) , = ( J A * ( k ) + ϵ ( k ) ) Δ q ( k ) + Δ χ d ( k + 1 ) , Δ e ( k + 1 ) = ( J A * ( k ) + ϵ ( k ) ) J ^ A ( k ) C k α * ( k ) e ( k ) α ( k ) Δ χ d ( k + 1 ) + Δ χ d ( k + 1 ) , ( J A * ( k ) + ϵ ( k ) ) N ^ ( k ) Δ z ( k ) .
In Assumption 2, it is established that if Δ J ^ A ( k ) = J ^ A ( k + 1 ) J ^ A ( k ) = 0 , then J ^ A ( k ) = J A * ( k ) ; thus, ϵ ( k ) = 0 . Therefore, ( J A * ( k ) + ϵ ( k ) ) J ^ A ( k ) = M k and ( J A * ( k ) + ϵ ( k ) ) N ^ ( k ) Δ z ( k ) = Z k , if Δ J ^ A ( k ) 0 , then M k J ^ A ( k ) J ^ A ( k ) = I and Z k J ^ A ( k ) N ^ ( k ) Δ z ( k ) = 0 , where M k encodes the uncertainty during the learning process and Z k introduces disturbances from the unlearned interaction of null-space velocities. Thus, M k = I and Z k = 0 after the learning process. If ρ ( J ^ A ( k ) ) = m , then M k is a positive definite matrix as it satisfies M k = M k T , m i , i > i j | m i , j |   i , and m i , i > 0   i . Therefore, during the learning process, (A6) becomes:
Δ e ( k + 1 ) = M k C k α * e ( k ) + M k C k α ( k ) I Δ χ d ( k + 1 ) Z k = M k C k α * e ( k ) + Ω k ,
where Ω k = M k C k α ( k ) I Δ χ d ( k + 1 ) Z k . Note that, after the learning process, i.e., when Δ J ^ A ( k ) = 0 , M k = I , Z k = 0 , and α ( k ) = C k 1 , Δ e ( k + 1 ) simply becomes:
Δ e ( k + 1 ) = C k α * e ( k ) + C k α ( k ) I Δ χ d ( k + 1 ) = a e ( k ) .
By substituting (A8) into (A4), the discrete change of the Lyapunov function becomes:
Δ V c ( k + 1 ) = a e ( k ) T e ( k ) + 1 2 a 2 e ( k ) T e ( k ) ,
where the negative term is dominant if a < 1 , which is always the case since a = α 1 α 1 + α 2 < 1 , for α 1 < α 2 .
Now, for the analysis of the learning phase, with α ( k ) = 1 , α * ( k ) = a and Ω k = M k C k I Δ χ d ( k + 1 ) Z k ; substitute (A7) into (A4) to get:
Δ V c ( k + 1 ) = [ M k C k a e ( k ) + Ω k ] T e ( k ) 1 2 [ M k C k a e ( k ) Ω k ] ,
expanding its terms:
Δ V c ( k + 1 ) = [ M k C k a e ( k ) ] T e ( k ) + 1 2 [ M k C k a e ( k ) ] T [ M k C k a e ( k ) Ω k ] + Ω k T e ( k ) 1 2 Ω k T [ M k C k a e ( k ) Ω k ] , = a C k M k e ( k ) T e ( k ) + a 2 2 M k M k C k C k e ( k ) T e ( k ) a C k M k Ω k T e ( k ) + Ω k T e ( k ) + 1 2 Ω k T Ω k .
Note that setting α 2 α 1 ensures dominance of the negative term against the quadratic positive one but yields a small a, which can make it less dominant against the rest of the opposing terms; thus, a balanced maximum value of a = 0.5 , namely when α 1 = α 2 , is recommended:
( 0.5 ) C k M k e ( k ) T e ( k ) > ( 0.125 ) M k M k C k C k e ( k ) T e ( k ) .
Moreover, ς i in C k should be chosen such that ς i > 1 J ^ A i ( k ) max 2 , guaranteeing that C k > C k T C k , further enforcing (A12). Since M k is positive definite during learning, its eigenvalues bounds are λ min ( M k ) λ i ( M k ) λ max ( M k ) , with λ min ( M k ) > 0 , λ max 1 for small ϵ . Also note that λ min ( M k ) = λ min ( M k ) = 1 after learning. Moreover, the bound of J ^ A ( k ) depends on ρ ( J ^ A ) = m and its regularization factor μ from (4), such that:
J ^ A ( k ) 1 σ min ( J ^ A ( k ) ) + μ ,
where σ min ( J ^ A ( k ) ) is the smallest singular value of J ^ A ( k ) . Thus, a well-conditioned M k is bounded as:
1 ϵ ( k ) J ^ A ( k ) λ min ( M k ) λ max ( M k ) 1 + ϵ ( k ) J ^ A ( k ) ,
which largely depends on η and μ from (4). Note that Ω k 0 when ϵ 0 , which means it is a vanishing term during the learning process, with M k being bounded by (A14) and C k being designed such that C k < C k T C k . Moreover, χ d ( k ) can be defined by a smooth continuous function, such that Δ χ d ( k + 1 ) C 1 , or be a measured signal χ d * ( k ) under Assumption 2, similar to χ ( k ) * . Under these considerations Δ χ d ( k + 1 ) < L is inherently bounded by the Lipchitz constant, in line with Assumption 1. The remaining term to analyze is Z k , which also depends on ϵ 0 , such that ( J A * ( k ) + ϵ ( k ) ) N ^ ( k ) 0 . Therefore, during learning, Z k introduces null-space velocities into the range-space with N ^ ( k ) = I J ^ A ( k ) J ^ A ( k ) , where J ^ A ( k ) satisfies (A13); therefore, N k also satisfies:
N ^ ( k ) 1 σ min ( J ^ A ( k ) ) σ min ( J ^ A ( k ) ) + μ ,
where μ must be the dominant term for a small σ min ( J ^ A ( k ) ) . The effect of Z k in the range-space is proportional to
Z k ( J A * ( k ) + ϵ ( k ) ) N ^ ( k ) Δ z ( k ) ,
where J A * ( k ) is bounded by (A13), ϵ ( k ) by (4), and N ^ ( k ) by (A15), whereas Δ z ( k ) defaults to zero if no explicit null-space objectives are specified, otherwise it can be set to minimize an objective function like (33); hence, Z k is bounded for a well-defined Δ z ( k ) . Also, triangle inequality leads to Ω k Γ k , where Γ k = λ max ( G ) Δ χ d ( k + 1 ) + Z k , with G = M k C k I , which can be used to summarize (A11) and its terms involving Ω k as follows:
| a C k M k Ω k T e ( k ) | a C k M k e ( k ) Ω k a C k M k e ( k ) Γ k ,
| Ω k T e ( k ) | e ( k ) Ω k e ( k ) Γ k ,
1 2 Ω k T Ω k = 1 2 Ω k 2 1 2 Γ 2 .
Finally, (A11) becomes:
Δ V c ( k + 1 ) A + B + C ,
with
A = a e ( k ) 2 [ λ min ( M k C k ) ] ,
B = a 2 2 e ( k ) 2 λ max 2 ( M k C k ) ,
C = Γ k ( a M k C k e ( k ) + e ( k ) ) + 1 2 Γ k 2 ,
where (A12) proves that A > B . Now let us prove that A > C . First, note that A scales quadratically with a e ( k ) 2 , while C scales linearly with Γ k e ( k ) and Γ k a M k C k e ( k ) , so for a large e ( k ) , A will dominate over C . Furthermore, M k C k e ( k ) λ max ( M k C k ) e ( k ) . Therefore:
C Γ k ( a λ max ( M k C k ) e ( k ) + e ( k ) ) + 1 2 Γ k 2 , Γ k e ( k ) ( a λ max ( M k C k ) + 1 ) + 1 2 Γ k 2 .
Then, for A > C to hold:
a e ( k ) 2 [ λ min ( M k C k ) ] > Γ k e ( k ) ( a λ max ( M k C k ) + 1 ) + 1 2 Γ k 2 ,
a e ( k ) [ λ min ( M k C k ) ] > Γ k ( a λ max ( M k C k ) + 1 ) + 1 2 Γ k 2 e ( k ) ,
where
a [ λ min ( M k C k ) ] > Γ k ( a λ max ( M k C k ) + 1 ) e ( k ) .
Thus, for a large e ( k ) , the discrete change of Lyapunov function remains strictly negative, Δ V c ( k + 1 ) < 0 . Otherwise, there is a minimum error threshold for the stability given by:
e ( k ) min > Γ k ( a λ max ( M k C k ) + 1 ) + 1 2 Γ k 2 a λ min ( M k C k ) ,
below which the strict negativity of Δ V c ( k + 1 ) < 0 cannot be guaranteed. For e ( k ) < e ( k ) min , stability holds if Δ V c ( k + 1 ) 0 , which ensures boundedness of the system, i.e., V c ( k + 1 ) V c ( k ) .
Finally, note that for e ( k ) to be small, it requires ϵ ( k ) 0 , such that J ^ A ( k ) = J A * ( k ) , as e ( k ) 0 is possible only through a proper mapping, J ^ A : R m R n , which means the learning phase is concluded and (A20) becomes (A9). This concludes the stability proof.

References

  1. Islam, R.U.; Iqbal, J.; Manzoor, S.; Khalid, A.; Khan, S. An autonomous image-guided robotic system simulating industrial applications. In Proceedings of the 2012 7th International Conference on System of Systems Engineering (SoSE), Genoa, Italy, 16–19 July 2012; IEEE: Piscataway Township, NJ, USA, 2012; pp. 344–349. [Google Scholar]
  2. Yu, Q.; Wang, G.; Hua, X.; Zhang, S.; Song, L.; Zhang, J.; Chen, K. Base position optimization for mobile painting robot manipulators with multiple constraints. Robot. Comput.-Integr. Manuf. 2018, 54, 56–64. [Google Scholar] [CrossRef]
  3. Gracia, L.; Solanes, J.E.; Munoz-Benavent, P.; Miro, J.V.; Perez-Vidal, C.; Tornero, J. Adaptive sliding mode control for robotic surface treatment using force feedback. Mechatronics 2018, 52, 102–118. [Google Scholar] [CrossRef]
  4. Saxena, A.; Kumar, J.; Sharma, K.; Roy, D. Controlling of Manipulator for Performing Advance Metal Welding. In Recent Innovations in Mechanical Engineering: Select Proceedings of ICRITDME 2020; Springer: Berlin/Heidelberg, Germany, 2022; pp. 41–48. [Google Scholar]
  5. Awan, Z.S.; Ali, K.; Iqbal, J.; Mehmood, A. Adaptive backstepping based sensor and actuator fault tolerant control of a manipulator. J. Electr. Eng. Technol. 2019, 14, 2497–2504. [Google Scholar] [CrossRef]
  6. Hou, Z.; Chi, R.; Gao, H. An overview of dynamic-linearization-based data-driven control and applications. IEEE Trans. Ind. Electron. 2016, 64, 4076–4090. [Google Scholar] [CrossRef]
  7. Treesatayapun, C. Data-driven fault-tolerant control with fuzzy-rules equivalent model for a class of unknown discrete-time MIMO systems and complex coupling. J. Comput. Sci. 2022, 63, 101827. [Google Scholar] [CrossRef]
  8. Treesatayapun, C. Fuzzy rules emulated discrete-time controller based on plant’s input–output association. J. Control Autom. Electr. Syst. 2019, 30, 902–910. [Google Scholar] [CrossRef]
  9. Campos-Torres, I.; Gómez, J.; Baltazar, A. Data-Driven Adaptive Force Control for a Novel Soft-Robot Based on Ultrasonic Atomization. In Proceedings of the Advances in Computational Intelligence: 21st Mexican International Conference on Artificial Intelligence, MICAI 2022, Monterrey, Mexico, 24–29 October 2022; Proceedings, Part II. Springer: Cham, Switzerland, 2022; pp. 279–290. [Google Scholar]
  10. Gao, S.; Zhao, D.; Yan, X.; Spurgeon, S.K. Model-Free Adaptive State Feedback Control for a Class of Nonlinear Systems. IEEE Trans. Autom. Sci. Eng. 2023, 21, 1824–1836. [Google Scholar] [CrossRef]
  11. Liu, T.; Hou, Z. Model-Free Adaptive Containment Control for Unknown Multi-Input Multi-Output Nonlinear MASs With Output Saturation. IEEE Trans. Circuits Syst. Regul. Pap. 2023, 70, 2156–2166. [Google Scholar] [CrossRef]
  12. Gómez, J.; Treesatayapun, C.; Morales, A. Data-driven identification and control based on optic tracking feedback for robotic systems. Int. J. Adv. Manuf. Technol. 2021, 113, 1485–1503. [Google Scholar] [CrossRef]
  13. Li, M.; Kang, R.; Branson, D.T.; Dai, J.S. Model-free control for continuum robots based on an adaptive Kalman filter. IEEE/ASME Trans. Mechatron. 2017, 23, 286–297. [Google Scholar] [CrossRef]
  14. Villalobos, J.; Sanchez, I.Y.; Martell, F. Singularity Analysis and Complete Methods to Compute the Inverse Kinematics for a 6-DOF UR/TM-Type Robot. Robotics 2022, 11, 137. [Google Scholar] [CrossRef]
  15. Abbes, M.; Poisson, G. Geometric Approach for Inverse Kinematics of the FANUC CRX Collaborative Robot. Robotics 2024, 13, 91. [Google Scholar] [CrossRef]
  16. Campa, R.; De La Torre, H. Pose control of robot manipulators using different orientation representations: A comparative review. In Proceedings of the 2009 American Control Conference, St. Louis, MO, USA, 10–12 June 2009; IEEE: Piscataway Township, NJ, USA, 2009; pp. 2855–2860. [Google Scholar]
  17. Bai, Q.; Shehata, M.; Nada, A. Review study of using Euler angles and Euler parameters in multibody modeling of spatial holonomic and non-holonomic systems. Int. J. Dyn. Control 2022, 10, 1707–1725. [Google Scholar] [CrossRef]
  18. Gonçalves, F.; Ribeiro, T.; Ribeiro, A.F.; Lopes, G.; Flores, P. A recursive algorithm for the forward kinematic analysis of robotic systems using euler angles. Robotics 2022, 11, 15. [Google Scholar] [CrossRef]
  19. Gómez, J.; Treesatayapun, C.; Morales, A. Multi-Inputs and Multi-Outputs equivalent model based on data driven controller for a robotic system. IFAC-PapersOnLine 2020, 53, 9784–9789. [Google Scholar] [CrossRef]
  20. Nguyen, H.T.; Cheah, C.C.; Toh, K.A. A data-driven iterative learning algorithm for robot kinematics approximation. In Proceedings of the 2019 IEEE/ASME International Conference on Advanced Intelligent Mechatronics (AIM), Hong Kong, China, 8–12 July 2019; IEEE: Piscataway Township, NJ, USA, 2019; pp. 1031–1036. [Google Scholar]
  21. Nguyen, H.T.; Cheah, C.C. Data-driven neural network-based learning for regression problems in robotics. In Proceedings of the IECON 2020 The 46th Annual Conference of the IEEE Industrial Electronics Society, Singapore, 18–21 October 2020; IEEE: Piscataway Township, NJ, USA, 2020; pp. 581–586. [Google Scholar]
  22. Wu, H.; Jin, S.; Yin, C.; Zheng, J.; Hou, Z. Model free adaptive predictive tracking control for robot manipulators with uncertain parameters. In Proceedings of the 2021 IEEE 10th Data Driven Control and Learning Systems Conference (DDCLS), Suzhou, China, 14–16 May 2021; IEEE: Piscataway Township, NJ, USA, 2021; pp. 1571–1576. [Google Scholar]
  23. Fan, J.; Jin, L.; Xie, Z.; Li, S.; Zheng, Y. Data-driven motion-force control scheme for redundant manipulators: A kinematic perspective. IEEE Trans. Ind. Informatics 2021, 18, 5338–5347. [Google Scholar] [CrossRef]
  24. Marconi, G.M.; Camoriano, R.; Rosasco, L.; Ciliberto, C. Structured prediction for CRiSP inverse kinematics learning with misspecified robot models. IEEE Robot. Autom. Lett. 2021, 6, 5650–5657. [Google Scholar] [CrossRef]
  25. Morales-Díaz, A.B.; Gómez-Casas, J.; Treesatayapun, C.; Muñiz-Valdez, C.R.; Galindo-Valdés, J.S.; Martínez-Villafañe, J.F. Data-Driven Adaptive Modelling and Control for a Class of Discrete-Time Robotic Systems Based on a Generalized Jacobian Matrix Initialization. Mathematics 2023, 11, 2555. [Google Scholar] [CrossRef]
  26. Lynch, K.M.; Park, F.C. Modern Robotics: Mechanics, Planning, and Control, 1st ed.; Cambridge University Press: New York, NY, USA, 2017. [Google Scholar]
  27. Hou, Z.; Jin, S. Data-driven model-free adaptive control for a class of MIMO nonlinear discrete-time systems. IEEE Trans. Neural Netw. 2011, 22, 2173–2188. [Google Scholar] [PubMed]
  28. Gómez, J.; Morales, A.; Treesatayapun, C.; Muñiz, R. Data-driven-modelling and Control for a Class of Discrete-Time Robotic System Using an Adaptive Tuning for Pseudo Jacobian Matrix Algorithm. In Proceedings of the Advances in Computational Intelligence: 21st Mexican International Conference on Artificial Intelligence, MICAI 2022, Monterrey, Mexico, 24–29 October 2022; Proceedings, Part II. Springer: Cham, Switzerland, 2022; pp. 291–302. [Google Scholar]
  29. Toro-Arcila, C.A.; Becerra, H.M.; Arechavaleta, G. Visual path following with obstacle avoidance for quadcopters in indoor environments. Control Eng. Pract. 2023, 135, 105493. [Google Scholar] [CrossRef]
Figure 1. Closed-loop configuration diagram.
Figure 1. Closed-loop configuration diagram.
Processes 12 02831 g001
Figure 2. 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: (a) UR5 simulation setup, (b) KY simulation setup, and (c) KYD simulation setup.
Figure 2. 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: (a) UR5 simulation setup, (b) KY simulation setup, and (c) KYD simulation setup.
Processes 12 02831 g002
Figure 3. End-effector user commands.
Figure 3. End-effector user commands.
Processes 12 02831 g003
Figure 4. End-effector pose errors.
Figure 4. End-effector pose errors.
Processes 12 02831 g004
Figure 5. End-effector velocities.
Figure 5. End-effector velocities.
Processes 12 02831 g005
Figure 6. Control signals.
Figure 6. Control signals.
Processes 12 02831 g006
Figure 7. End-effector user commands.
Figure 7. End-effector user commands.
Processes 12 02831 g007
Figure 8. End-effector pose errors.
Figure 8. End-effector pose errors.
Processes 12 02831 g008
Figure 9. End-effector velocities.
Figure 9. End-effector velocities.
Processes 12 02831 g009
Figure 10. Control signals.
Figure 10. Control signals.
Processes 12 02831 g010
Figure 11. Commands for the left tip.
Figure 11. Commands for the left tip.
Processes 12 02831 g011
Figure 12. Commands for the right tip.
Figure 12. Commands for the right tip.
Processes 12 02831 g012
Figure 13. Pose errors of the left tip.
Figure 13. Pose errors of the left tip.
Processes 12 02831 g013
Figure 14. Pose errors of the right tip.
Figure 14. Pose errors of the right tip.
Processes 12 02831 g014
Figure 15. Left end-effector velocities.
Figure 15. Left end-effector velocities.
Processes 12 02831 g015
Figure 16. Right end-effector velocities.
Figure 16. Right end-effector velocities.
Processes 12 02831 g016
Figure 17. Control signals.
Figure 17. Control signals.
Processes 12 02831 g017
Figure 18. CoppeliaSim simulation environment: the KY’s end-effector performs a trajectory tracking task defined in the ( x , y ) plane, with a constant height z, while keeping the initial orientation unchanged—the goal is to perform a polishing operation under the surface of the disk. (a) Simulation setup home position. (b) KY under DDMC scheme. (c) KY under classic inverse-kinematics control scheme.
Figure 18. CoppeliaSim simulation environment: the KY’s end-effector performs a trajectory tracking task defined in the ( x , y ) plane, with a constant height z, while keeping the initial orientation unchanged—the goal is to perform a polishing operation under the surface of the disk. (a) Simulation setup home position. (b) KY under DDMC scheme. (c) KY under classic inverse-kinematics control scheme.
Processes 12 02831 g018
Figure 19. Pose control performance indexes while performing a cyclic routine. (a) Performance indexes under DDCM scheme. (b) Performance indexes under classic inverse-kinematics control scheme.
Figure 19. Pose control performance indexes while performing a cyclic routine. (a) Performance indexes under DDCM scheme. (b) Performance indexes under classic inverse-kinematics control scheme.
Processes 12 02831 g019
Table 1. Parameters setting for the robot controller.
Table 1. Parameters setting for the robot controller.
Parametersxyz θ x θ y θ z
α 1 i 2.52.52.52.52.52.5
α 2 i 2.52.52.52.52.52.5
ς i 0.650.650.650.650.650.65
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Goméz-Casas, J.; Toro-Arcila, C.A.; Rodríguez-Rosales, N.A.; Obregón-Flores, J.; Ortíz-Ramos, D.E.; Martínez-Villafañe, J.F.; Gómez-Casas, O. Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot. Processes 2024, 12, 2831. https://doi.org/10.3390/pr12122831

AMA Style

Goméz-Casas J, Toro-Arcila CA, Rodríguez-Rosales NA, Obregón-Flores J, Ortíz-Ramos DE, Martínez-Villafañe JF, Gómez-Casas O. Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot. Processes. 2024; 12(12):2831. https://doi.org/10.3390/pr12122831

Chicago/Turabian Style

Goméz-Casas, Josué, Carlos A. Toro-Arcila, Nelly Abigaíl Rodríguez-Rosales, Jonathan Obregón-Flores, Daniela E. Ortíz-Ramos, Jesús Fernando Martínez-Villafañe, and Oziel Gómez-Casas. 2024. "Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot" Processes 12, no. 12: 2831. https://doi.org/10.3390/pr12122831

APA Style

Goméz-Casas, J., Toro-Arcila, C. A., Rodríguez-Rosales, N. A., Obregón-Flores, J., Ortíz-Ramos, D. E., Martínez-Villafañe, J. F., & Gómez-Casas, O. (2024). Data-Driven Kinematic Model for the End-Effector Pose Control of a Manipulator Robot. Processes, 12(12), 2831. https://doi.org/10.3390/pr12122831

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop