Next Article in Journal
Non-Selective Reduction of P-Stereogenic Phosphinoylacetic Acid Esters and 3-Phosphorylated Coumarins to Phosphino-Boranes: Discovery of Unexpected 2,3-Dihydrobenzofuran Derivative
Previous Article in Journal
Vitali Theorems for Varying Measures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

State Reconstruction of Remote Robotic System Using Delayed Output and Torque Estimation

School of Microelectronics and Control Engineering, Changzhou University, Changzhou 213164, China
*
Author to whom correspondence should be addressed.
Symmetry 2024, 16(8), 975; https://doi.org/10.3390/sym16080975
Submission received: 9 July 2024 / Revised: 23 July 2024 / Accepted: 26 July 2024 / Published: 31 July 2024

Abstract

:
The state reconstruction problem for a remote robotic system with communication delay is investigated in this paper. A symmetric estimation scheme is proposed based on both input and output observers. First, a sliding-mode disturbance observer is designed to estimate the input torque of the robot on the remote side. Then, by using the received output and torque estimation subject to time delay, a novel predictive observer is proposed on the local side to reconstruct the real-time joint angles and velocities. Based on a Lyapunov approach, sufficient conditions are obtained to make sure that the estimation errors can converge exponentially to bounded regions through selecting proper observer gains. Simulation studies are conducted to verify the effectiveness of the proposed strategy. The estimation error of state reconstruction is decreased by 14% in terms of the integral square error when compared to the standard high-gain predictor, and the simulation results demonstrate the effectiveness of force estimation under disturbances and model uncertainties.

1. Introduction

Time delays are common issues in the application of remote robotic systems with long communication distances, such as space robots, underwater robots, and teleoperation systems [1,2,3,4,5]. Even a small delay can degrade the performance and stability of the robot control system [6]. A number of studies have investigated the control schemes for robotic systems with uncertainties and disturbances, such as data-based PID control [7] and fuzzy control [8]. Meanwhile, various techniques have focused on the controller design for robot systems with input delay, such as predictive controllers [9,10], adaptive controllers [11], and neural network controllers [12], which ensure the stability of the remote systems. On the other hand, however, the delay feedback signal on the local side will cause the operator to be unable to grasp the remote working status in time and thus be unable to issue the correct control instructions. Thus, it is necessary to reconstruct the state of remote robotic systems by using delayed feedback.
For nonlinear systems with output delay, high-gain observers/predictors are designated in [13,14,15] to estimate the system’s state based on delayed measurements. And this approach is extended to cases of time-varying delays in [16,17] where Lyapunov methods are applied to analyze the convergence of estimation errors. In [18,19,20], continuous-discrete time observers are applied as underlying systems by using sampled and delayed outputs. Then, cascade observers are proposed in [21,22,23], which are composed of a number of sub-observers in chains, such that the large time delays can be divided into small segmentations. Additionally, an adaptive observer is designed in [24] for heat diffusion-reaction partial differential equations with sensor delay. In [25], Immersion and Invariance techniques are used to design observers for general nonlinear systems when the output measurements are subject to constant time delays. However, the input of the system is assumed to be known in these methods. For remote robotic systems, the input torque is hard to measure and often suffers from disturbances. In addition, only the delayed estimation of input can be applied to the observer design due to the time delay.
The issues of force/torque estimations for robot control systems have been addressed in many previous studies [26,27,28]. For example, in [29,30], force estimations are applied to the controller designs for telerobotics, which consist of symmetric leader and follower robots, such that the operational performance is enhanced under time delay. For robotic systems without force sensors, the disturbance observer has been widely applied to estimate the external input of robotic systems due to its simple structure and effectiveness [31]. In [32], a globally exponentially stable observer for joint torque information is constructed using only position measurement. In [33], null-space impedance control of a kinematically redundant robot is investigated based on a disturbance observer. Recently, sliding-mode approaches have been used by disturbance observers to ensure asymptotic convergence of estimation errors [34,35,36]. However, state reconstruction based on delayed torque estimation remains an open problem.
In this paper, a symmetric state estimation scheme is proposed for a remote robot system with time delay. First, a sliding-mode disturbance observer is designed to estimate input torque on the remote (robot) side. Then, based on the delayed torque estimation and the output received on the local side, a novel predictive observer is proposed to reconstruct the real-time joint angles and velocities of the robotic system. The main novelties of this work are summarized as follows:
  • The robustness to disturbances and uncertainties of the proposed sliding-mode observer is improved by applying the super-twisting algorithm [37]. The estimation error of input torque converges in finite time to 0.
  • A novel predictive algorithm is proposed to design the state observer based on the Euler–Lagrange model and the delayed output. Through using Lyapunov approaches, sufficient conditions are obtained to make sure that the estimation errors can converge exponentially to bounded regions through the selection of proper observer gains.
This paper is organized as follows: The mathematical model and problem formulation are presented in Section 2. The main results, including the torque estimation and state reconstruction, are shown in Section 3. Numerical simulations are performed in Section 4 to confirm the effectiveness of the method. Finally, the results are discussed in Section 5.

2. Problem Formulation

Consider an n-degree-of-freedom (DOF) robotic system described by the following Euler–Lagrange equation:
M ( q ) q ¨ + C ( q , q ˙ ) q ˙ + G ( q ) = τ
where q n and q ˙ n are the joint angles and velocities of the robots, respectively, τ represents the unknown input torque imposed by controller and external disturbances, M ( q ) n × n is the symmetric and positive-definite inertia matrix, C ( q , q ˙ ) n × n is the matrix of the centripetal and Coriolis torques, and G ( q ) n × n represents the vector of the gravitational forces.
As shown in Figure 1, the control inputs (such as desired trajectories) are transferred to the remote robot through a communication channel. There is a communication delay d between the remote robot and the local control side. The objective is to obtain an estimation τ ^ of the input torque τ on the remote side and then reconstruct the real-time joint angles and velocities on the local side by using the delayed signals, including τ ^ ( t d ) , q ( t d ) , and q ˙ ( t d ) . The following assumptions are useful in the design:
Assumption 1.
The system states  q   and q ˙  are bounded.
Assumption 2.
The input torque  τ  and its derivative are bounded. Let η k 1 M 1 ( q ) τ ; there are positive constants ϕ  and v such that
| η ˙ + k λ 5 η | i < ϕ
t d t η ( σ ) d σ 2 v
where i represents the absolute value of the i-th element of the vector for i = 1 , , n ; denotes the standard Euclidean norm; λ 5 and k are constant parameters which will be specified later.
Remark 1.
The state and input boundaries are common assumptions in many works of observer design [21,22]. In addition, the joint angle, velocity, and input torque are always bounded in practical robotic applications. Noticing that the matrix M 1 ( q )  is bounded when q  is bounded [34], inequalities (2)–(3) are reasonable as long as τ  is bounded.

3. Main Results

3.1. Force Estimation Based on Sliding-Mode Disturbance Observer

Let q 1 = q , q 2 = q ˙ . According to dynamic model (1), there is
q ˙ 2 = M 1 ( q 1 ) [ C ( q 1 , q 2 ) q 2 + G ( q 1 ) τ ] .
The sliding-mode disturbance observer is designed to estimate the input torque τ by using the super-twisting algorithm [37] as follows:
q ^ ˙ 2 = M 1 ( q 1 ) [ C ( q 1 , q 2 ) q 2 + G ( q 1 ) ] + λ 1 q 2 q ^ 2 sgn ( q 2 q ^ 2 ) + λ 2 ( q 2 q ^ 2 ) + k η ^
η ^ ˙ = λ 3 sgn ( q 2 q ^ 2 ) + λ 4 ( q 2 q ^ 2 ) λ 5 η ^
where λ 1 , λ 2 , λ 3 , λ 4 , λ 5 , and k are positive constants, sgn() represents the sign function, and η ^ is an estimation of η k 1 M 1 ( q ) τ . Thus, the estimated input torque can be obtained as
τ ^ = k M ( q ) η ^
The estimation errors of the observer are defined as e 1 = q 2 q ^ 2 and e 2 = η η ^ . It can be calculated from (4)–(6) that
e ˙ 1 = λ 1 q 2 q ^ 2 sgn ( q 2 q ^ 2 ) λ 2 ( q 2 q ^ 2 ) + k e ˙ 2
e ˙ 2 = k λ 3 sgn ( q 2 q ^ 2 ) k λ 4 ( q 2 q ^ 2 ) k λ 5 e 2 + η ˙ + k λ 5 η
Theorem 1.
Consider dynamic systems of estimation error (8)–(9) under Assumptions 1 and 2. The gains λ 1 , λ 2 , λ 3 , λ 4 , λ 5 , and  k  can be selected with high enough values such that  e 1  and  e 2  converge in finite time to the origin  e 1 = 0  and  e 2 = 0 .
Proof of Theorem 1.
Consider a Lyapunov function candidate [34]:
V = i = 1 n V i
V i = 1 2 ξ i T Q ξ i
where ξ i T = e 1 i 1 / 2 sgn ( e 1 i ) e 1 i e 2 i , and Q = 4 k λ 3 + λ 1 2 λ 1 λ 2 λ 1 λ 1 λ 2 2 k λ 4 + λ 2 2 λ 2 λ 1 λ 2 2 .
Taking the time derivative of V i yields
V ˙ i = 1 e 1 i 1 / 2 ξ i T Ω 1 ξ i ξ i T Ω 2 ξ i λ 1 ( η ˙ + k λ 5 η ) e 1 i 1 / 2 sgn ( e 1 i ) λ 2 ( η ˙ + k λ 5 η ) e 1
where Ω 1 = λ 1 2 2 k λ 3 + λ 1 2 0 λ 1 0 2 k λ 4 + 5 λ 2 2 3 λ 2 λ 1 3 λ 2 1 , Ω 2 = λ 2 k λ 3 + 2 λ 1 2 0 1 2 k λ 1 λ 5 0 k λ 4 + λ 2 2 λ 2 1 2 k λ 1 λ 5 λ 2 1 .
By applying condition (2), there is
λ 1 ( η ˙ + k λ 5 η ) e 1 i 1 / 2 sgn ( e 1 i ) λ 2 ( η ˙ + k λ 5 η ) e 1 < λ 1 δ e 1 i 1 / 2 + λ 2 δ e 1
Thus, the upper bound of V ˙ is determined as
V ˙ < i = 1 n ( 1 e 1 i 1 / 2 ξ i T Ω 3 ξ i + ξ i T Ω 4 ξ i )
where Ω 3 = λ 1 2 2 k λ 3 + λ 1 2 λ 1 ϕ 0 λ 1 0 2 k λ 4 + 5 λ 2 2 3 λ 2 λ 1 3 λ 2 1 , Ω 4 = λ 2 k λ 3 + 2 λ 1 2 λ 2 ϕ 0 1 2 k λ 1 λ 5 0 k λ 4 + λ 2 2 λ 2 1 2 k λ 1 λ 5 λ 2 1 .
If the gains are selected with high enough values such that Q , Ω 3 , and Ω 4 are positive-definite matrices, then there are positive constants c 1 and c 2 such that
V ˙ < c 1 V 1 2 c 2 V
Based on the comparison lemma, it can be concluded that e 1 and e 2 converge to 0 in finite time, which is less than T = 2 V 1 / 2 c 1 . The proof of Theorem 1 is completed. □
Remark 2.
It should be pointed out that the time derivative of  V i  in (12) is similar but not the same as the results in [37] due to the existence of an additional term,  k λ 5 e 2 , in error dynamic function (9). From the definition of  Ω 3  and  Ω 4 , if  λ 1  and  λ 3  are sufficiently large relative to  ϕ , then there are proper  λ 2 , λ 4 , λ 5 , and k such that  Ω 3  and  Ω 4  are positive-definite.

3.2. State Reconstruction Using Delayed Output and Torque Estimation

By using the delayed outputs q 1 d ( t ) and q 2 d ( t ) and the estimation η ^ d ( t ) = η ^ ( t d ) , the original state variables of system (1) are reconstructed by the following predictive observer:
x ^ ˙ 1 ( t ) = x ^ 2 ( t ) , x ^ ˙ 2 ( t ) = M 1 ( x ^ 1 ( t ) ) [ C ( x ^ 1 ( t ) , x ^ 2 ( t ) ) x ^ 2 ( t ) + G ( ( x ^ 1 ( t ) ) ] + k η ^ d ( t ) + K p r d ( t ) , r d ( t ) = q 2 d ( t ) x ^ 2 d ( t ) + α ( q 1 d ( t ) x ^ 1 d ( t ) ) t d t K p r d ( σ ) + k η ^ d ( σ ) d σ
where x ^ 1 ( t ) and x ^ 2 ( t ) represent the estimations of q 1 ( t ) and q 2 ( t ) , respectively, r d ( t ) is an auxiliary error variable, and K p is the designed constant gain, which can be expanded as K p = k a + k b + k c , x ^ 1 d ( t ) = q ^ 1 ( t d ) , and x ^ 2 d ( t ) = q ^ 2 ( t d ) .
Denote the estimation errors as follows:
ε 1 ( t ) = q 1 ( t ) x ^ 1 ( t )
ε 2 ( t ) = q 2 ( t ) x ^ 2 ( t )
θ ( t ) = t d t K p r ( σ ) d σ
r ( t ) = ε 2 ( t ) + α ε 1 ( t ) θ ( t ) t d t k η ( σ ) d σ
It should be noticed that r ( t d ) = r d ( t ) . By taking the time derivative of r ( t ) , we obtain
r ˙ = q ˙ 2 q ^ ˙ 2 + α ε 2 + ( K p r d + k η d ) ( K p r + k η ) = x ˙ 2 i + α ε 2 g ( x ^ 1 i , x ^ 2 i ) M ^ i 1 ( x ^ 1 i ) u ^ i + k e 2 d k η ^ K p i r i = N ( q 1 , q 2 , x ^ 1 i , x ^ 2 i ) ε 1 + k ( e 2 d + e 2 ) K p r
e 2 = η η ^ was introduced in (9), and the auxiliary vectors N ( q 1 , q 2 , x ^ 1 i , x ^ 2 i ) are defined as
N ( q 1 , q 2 , x ^ 1 i , x ^ 2 i ) = M 1 ( q 1 ) [ C ( q 1 , q 2 ) q 2 + G ( q 1 ) ]   + M 1 ( x ^ 1 ( t ) ) [ C ( x ^ 1 ( t ) , x ^ 2 ( t ) ) x ^ 2 ( t ) + G ( x ^ 1 ( t ) ) ] + α ε 2 + ε 1
Similar to [38], the upper bound of N can be derived by applying the Mean Value Theorem as follows:
| | N | | ρ ( | | z | | ) | | z | |
where z = ε 1 T θ T r T T ; ρ ( ) is a positive invertible and a non-decreasing function.
Theorem 2.
Consider nonlinear system (1) and observer (16) under Assumptions 1 and 2. The estimation error z(t) convergences to bounded regions in the sense that
z ( t ) μ 0 e μ 1 t + μ 2
where μ 0 , μ 1 , μ 2 denote constants, provided that the following sufficient conditions are satisfied simultaneously:
d < 2 κ δ 3 K p 2 ,   k a > δ 2 + ω + κ d ,   k b > ρ 2 ( | | z ( 0 ) | | ) 4 c 1
where κ , δ , ω , c 1 are positive constants which will be specified later.
Proof of Theorem 2.
Consider the following Lyapunov function candidate:
V p = 1 2 ε 1 T ε 1 + 1 2 r T r + 1 2 K p θ T θ + Q + P ,
where Q = κ t d t s t | | r ( σ ) | | 2 d σ d s κ d t d t | | r ( σ ) | | 2 d σ , and P = ω t d t | | r ( σ ) | | 2 d σ .
Let y = ε 1 T r T θ T Q P T . It yields
s 1 | | y | | 2 V p ( y ) s 2 | | y | | 2 ,
where λ 1 = m i n 1 2 , 1 2 K p ,   λ 2 = m a x 1 , 1 2 K p .
Taking the time derivative of V, one obtains
V ˙ p = ε 1 T ( r α ε 1 + θ + t d t k η ^ ( σ ) d σ ) + r T ( N ε 1 + k ( e 2 d + e 2 ) K p r ) + θ T ( r d r ) + κ d | | r | | 2 κ t d t | | r ( σ ) | | 2 d σ + ω ( | | r | | 2 | | r d | | 2 ) .
The Cauchy–Schwarz inequality can be used to derive | θ | | 2 d K p 2 t d t | | r ( σ ) | | 2 d σ . Then, by utilizing Young’s inequality (3) and (23), the upper bound of V ˙ p is determined as
V ˙ i ( α δ ) | | ε 1 | | 2 + k 2 δ v + | | r | | ρ ( | | z | | ) | | z | | ( k b + k c ) | | r i | | 2 ( k a δ 2 ω κ d ) | | r | | 2 + k | | r | | | | e 2 d + e 2 | | ( ω δ 2 ) | | r d | | 2 κ d K p 2 3 2 δ γ 1 2 γ 2 2 | | θ | | 2 d ( γ 1 2 + γ 2 2 ) t d t | | r ( σ ) | | 2 d σ .   ( α δ ) | | ε 1 | | 2 ( k a δ 2 ω κ d ) | | r | | 2 κ d K p 2 3 2 δ γ 1 2 γ 2 2 | | θ | | 2 γ 1 2 κ Q d γ 2 2 ω P + ρ 2 ( | | z | | ) 4 k b | | z | | 2 + δ 2 k v + k | | e 2 d + e 2 | | 2 4 k c
By completing the squares based on the definition of z and y, the expression in (29) reduces to
V ˙ p c 1 ρ 2 ( | | z | | ) 4 k b | | z | | 2 γ 1 2 κ Q d γ 2 2 ω P + k 2 δ v + k | | e 2 d + e 2 | | 2 4 k c c 2 λ 2 V p + k 2 δ v + k | | e 2 d + e 2 | | 2 4 k c
where c 1 = min α δ 2 ,   k a δ 2 ω κ d ,   κ d K p 2 3 2 δ γ 1 2 γ 2 2 ,   c 2 = min c 1 ρ 2 ( | | z | | ) 4 k b , γ 1 2 κ , d γ 2 2 ω .
Because the convergence of e 2 was proved in Theorem 1, provided the conditions in (25) are satisfied, by using the comparison lemma, we can conclude that
V p V p ( y ( 0 ) ) e c 2 s 2 t + s 2 c 2 ( k 2 δ v + k | | e 2 d ( 0 ) + e 2 ( 0 ) | | 2 4 k c )
Then, in view of the definition of V p in (26), there exist proper constants μ 0 , μ 1 , μ 2 such that the result in (24) can be obtained. The proof of Theorem 2 is completed. □
From (31), since e 2 ( t ) converges in finite time to 0, the observer errors ε 1 ( t ) , r ( t ) , and θ ( t ) can converge to bounded regions which are determined by v and k. In view of the definition of r ( t ) in (20), ε 2 ( t ) is also bounded. Thus, we should apply a small value of k on the premise of ensuring the accuracy of torque estimation, in order to reduce the state reconstruction error. Additionally, from condition (25), the allowable delay by the observer is limited. One possible solution is to use several observers in cascade to deal with large time delays [39].
Remark 3.
The delayed auxiliary variable  η ^ d  is used directly in observer (16) in order to simplify the calculations. If just  τ ^ d  can be received on the local side, we should use  M 1 ( q ^ ) τ ^ d  instead of  k η ^ d . Under the circumstances, similar stability results of estimation errors can be obtained. Although the time delay is assumed to be constant in this paper, the proposed observer design approach can be straightforwardly applied to differentiable time-varying delays with d ˙ ( t ) < 1 [15,40].
Remark 4.
The proposed state reconstruction method can be applied to telerobotic systems with symmetric or asymmetric communication delays [29,30,41]. By using two symmetric estimation schemes on both the leader and follower sides, the time delay can be compensated through applying the reconstructed states of robots.

4. Simulations

To illustrate the performance of the proposed method, numerical simulations are performed on a 2-joint robot system with the following dynamics [41]:
M ( q ) = M 11 M 12 M 21 M 22 , C ( q , q ˙ ) = C 11 C 12 C 21 C 22
M 11 = ( 2 l a cos q b + l b ) l b m b + ( m a + m b ) l a 2 , M 12 = M 21 = l b 2 m b + l a l b m b cos q b , M 22 = l b 2 m b , C 11 = l a l b m b q ˙ b sin ( q b ) , C 12 = l a l b m b ( q ˙ a + q ˙ b ) sin ( q b ) , C 21 = l a l b m b q ˙ a sin ( q b ) , C 22 = 0 ,
G ( q ) = m a + m b g L a sin q a + m b g l b sin q a + q b m b g L b sin q a + q b ,
where the two joints are labeled as a and b, respectively, and q = q a T q b T T . The parameters are given as m a = 1   kg ,   m b = 0.5   kg ,   l a = 0.5   m , and l b = 0.5   m ; the initial states of the robot are given as q ( 0 ) = [ 0.1 0.1 ] rad and q ˙ ( 0 ) = [ 0 0 ] rad / s ; and the initial estimations are q ^ 1 ( 0 ) = x ^ 1 ( 0 ) = [ 0.1 0.1 ] rad and q ^ 2 ( 0 ) = x ^ 2 ( 0 ) = [ 0 0 ] rad / s . The input torque τ = τ a T τ b T T is a PD controller designed as τ = 30 ( q r q 1 ) + 10 ( q ˙ r q 2 ) , while the reference trajectory q r = q r a T q r b T T is set as q r a = 0.02 t   rad , 0 t 60   s 1.2   rad , t 65   s and q r b = 0.5 sin ( 0.2 t ) rad . The time delay d = 2   s , and the gain parameters are designed as λ 1 = 0.3 , λ 2 = 0.3 , λ 3 = 0.3 , λ 4 = 0.5 , λ 5 = 0.5 , k = 0.1 , K p = 20 , and α = 0.4 .
First, the torque estimation results of two joints (a and b) are shown in Figure 2. It can be seen that the estimations τ ^ a and τ ^ b coincide with the original torques τ a and τ b despite the jitter that happened due to the nature of the sliding-mode algorithm, which indicates the effectiveness of the sliding-mode disturbance observer. Then, the state reconstruction results of joint angles and velocities are given in Figure 3 and Figure 4, respectively. As can be seen, the good performance of the proposed predictive observer is realized, where the estimated states x ^ 1 and x ^ 2 are in correspondence with q 1 and q 2 .
Next, the proposed method is tested in different situations with disturbances and model uncertainties, respectively. The disturbance input is set as a uniform random number varying from 0.02 to 0.02   N m , and the nominal parameters used in observers (5) and (16) are given as m a = 0.5   kg ,   m b = 0.3   kg ,   l a = 0.4   m , and l b = 0.4   m , which are different from the real ones of the robot. The estimation errors e 2 under three different situations are compared in Figure 5, where the lines labeled “ideal” represent the results of the first simulation. It can be seen that the errors are more sensitive to disturbances than uncertainties because of the application of the sliding-mode approach. The strong robustness to model uncertainties of the proposed torque estimation method is verified. Moreover, the estimation errors of joint angles and velocities are shown in Figure 6 and Figure 7, which indicate that the proposed method can reconstruct the system state precisely when disturbances and model uncertainties exist.
Then, by setting the communication delay as d = 1 , 2 , 3   s , respectively, the reconstruction errors of joint angles are compared in Figure 8. It is obvious that the estimation error increases with the growth of the delay.
To verify the effectiveness of the proposed method, the state reconstruct results are compared with the cases without torque estimation and using the high-gain predictor [15]. First, when the parameters are set as d = 2   s and K p = 20 , the estimation results of the high-gain predictor are not convergent. The advantages of the proposed method in dealing with large time delays are shown. Thus, the parameters are adjusted to d = 1   s and K p = 1 for all three methods, and the estimation errors of joint angles are compared in Figure 9. It can be seen that the errors without using force estimation are higher than in the other two methods, and the results of the high-gain predictor are a bit of a wobble at the beginning. In addition, the integral square errors (ISEs) of ε 1 are presented and compared in Table 1, which are defined as 0 100 ε 1 ( t ) 2 d t . The results show that the estimation accuracy is improved by using the torque estimation, and the ISE of proposed method is reduced by 14% compared with the conventional high-gain predictor. The accuracy and rapidity of the proposed method in this paper are verified.

5. Conclusions

A predictive observer has been proposed to reconstruct the real-time state of remote robotic systems with time delay, while the input torque has been estimated by a sliding-mode disturbance observer. A comprehensive analysis has been carried out to investigate the fundamental properties of the proposed method. The effectiveness and robustness to disturbances and uncertainties have been verified by simulations. Future work will be focused on the observer and adaptive controller design [42] for symmetric or asymmetric telerobotics in more complex communication networks.

Author Contributions

Conceptualization, S.S.; Formal analysis, S.S. and L.J.; Funding acquisition, S.S.; Methodology, S.S.; Software, S.S.; Supervision, L.J.; Writing—original draft, S.S.; Writing—review and editing, L.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Natural Science Foundation of China under Grant No. 62303074 and by the Changzhou Sci&Tech Program under Grant No. CJ20235021.

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.

References

  1. Darvish, K.; Penco, L.; Ramos, J.; Cisneros, R.; Pratt, J.; Yoshida, E.; Ivaldi, S.; Pucci, D. Teleoperation of humanoid robots: A survey. IEEE Trans. Robot. 2023, 39, 1706–1727. [Google Scholar] [CrossRef]
  2. Mo, L.; Zhang, D.; Fu, X.; Dou, J.; Li, K.; Bai, L.; Bai, Y.; Zhang, Q.; Zhao, X. A multidirectional external perception soft actuator based on flexible optical waveguide for underwater teleoperation. Adv. Intell. Syst. 2023, 5, 2300029. [Google Scholar] [CrossRef]
  3. Qin, H.; Yang, H.; Sun, Y.; Feng, L. Fixed-time stable bilateral teleoperation of underwater manipulator using prescribed performance terminal sliding surfaces. J. Frankl. Inst. 2023, 360, 3280–3306. [Google Scholar] [CrossRef]
  4. Pacchierotti, C.; Prattichizzo, D. Cutaneous/tactile haptic feedback in robotic teleoperation: Motivation, survey, and perspectives. IEEE Trans. Robot. 2023, 40, 978–998. [Google Scholar] [CrossRef]
  5. Yan, J.; Gao, T.; Yang, X.; Chen, C.; Guan, X. Teleoperation Control of Autonomous Underwater Vehicle Toward Human on the Loop: Needs, Analyses, and Solutions. IEEE Syst. Man Cybern. Mag. 2024, 10, 2–13. [Google Scholar] [CrossRef]
  6. Roy, S.; Lee, J.; Baldi, S. A new adaptive-robust design for time delay control under state-dependent stability condition. IEEE Trans. Control Syst. Technol. 2020, 29, 420–427. [Google Scholar] [CrossRef]
  7. Ahmad, M.A.; Ishak, H.; Nasir, A.N.K.; Ghani, N.A. Data-based PID control of flexible joint robot using adaptive safe experimentation dynamics algorithm. Bull. Electr. Eng. Inform. 2021, 10, 79–85. [Google Scholar] [CrossRef]
  8. Li, J.; Pi, Y. Fuzzy time delay algorithms for position control of soft robot actuated by shape memory alloy. Int. J. Control. Autom. Syst. 2021, 19, 2203–2212. [Google Scholar] [CrossRef]
  9. González Sorribes, A.; García Gil, P. A novel observer-predictor control for uncertain systems with unknown time-varying input and output delays. Int. J. Control 2021, 94, 1630–1640. [Google Scholar] [CrossRef]
  10. Sharma, N.; Bhasin, S.; Wang, Q.; Dixon, W.E. Predictor-based control for an uncertain Euler–Lagrange system with input delay. Automatica 2011, 47, 2332–2342. [Google Scholar] [CrossRef]
  11. Keighobadi, J.; Fateh, M.M.; Xu, B. Adaptive fuzzy voltage-based backstepping tracking control for uncertain robotic manipulators subject to partial state constraints and input delay. Nonlinear Dyn. 2020, 100, 2609–2634. [Google Scholar] [CrossRef]
  12. Sun, W.; Wu, Y.; Lv, X. Adaptive neural network control for full-state constrained robotic manipulator with actuator saturation and time-varying delays. IEEE Trans. Neural Netw. Learn. Syst. 2021, 33, 3331–3342. [Google Scholar] [CrossRef]
  13. Adil, A.; Zemouche, A.; Hamaz, A.; Laleg-Kirati, T.; N’doye, I.; Bedouhene, F. High-gain observer design for nonlinear systems with delayed outputs. IFAC-PapersOnLine 2020, 53, 5057–5062. [Google Scholar] [CrossRef]
  14. Adil, A.; Hamaz, A.; N’doye, I.; Zemouche, A.; Laleg-Kirati, T.-M.; Bedouhene, F. On high-gain observer design for nonlinear systems with delayed output measurements. Automatica 2022, 141, 110281. [Google Scholar] [CrossRef]
  15. Ahmed-Ali, T.; Cherrier, E.; Lamnabhi-Lagarrigue, F. Cascade high gain predictors for a class of nonlinear systems. IEEE Trans. Autom. Control 2011, 57, 221–226. [Google Scholar] [CrossRef]
  16. Targui, B.; Hernández-González, O.; Astorga-Zaragoza, C.M.; Guerrero-Ramírez, G.-V.; Guerrero-Sánchez, M.-E. A new observer design for systems in presence of time-varying delayed output measurements. Int. J. Control. Autom. Syst. 2019, 17, 117–125. [Google Scholar] [CrossRef]
  17. Cacace, F.; Germani, A.; Manes, C. An observer for a class of nonlinear systems with time varying observation delay. Syst. Control Lett. 2010, 59, 305–312. [Google Scholar] [CrossRef]
  18. Tréangle, C.; Farza, M.; M’Saad, M. Observer design for a class of disturbed nonlinear systems with time-varying delayed outputs using mixed time-continuous and sampled measurements. Automatica 2019, 107, 231–240. [Google Scholar] [CrossRef]
  19. Ramírez-Rasgado, F.; Astorga-Zaragoza, C.M.; Hernández-González, O.; Guerrero-Sanchez, M.-E.; Osorio-Gordillo, G.-L.; Reyes-Reyes, J. Observer synthesis for uncertain nonlinear systems with nonuniformly sampled and delayed output. IEEE Syst. J. 2021, 16, 4532–4540. [Google Scholar] [CrossRef]
  20. Ammeh, L.; Giri, F.; Ahmed-Ali, T.; El Fadil, H. Sampled-data based observer design for nonlinear systems with output distributed delay. Int. J. Control 2021, 94, 1870–1879. [Google Scholar] [CrossRef]
  21. Farza, M.; Hernández-González, O.; Ménard, T.; Targui, B.; M’saad, M.; Astorga-Zaragoza, C.-M. Cascade observer design for a class of uncertain nonlinear systems with delayed outputs. Automatica 2018, 89, 125–134. [Google Scholar] [CrossRef]
  22. Hernández-González, O.; Farza, M.; Menard, T.; Targui, B.; M’saad, M.; Astorga-Zaragoza, C. A cascade observer for a class of MIMO non uniformly observable systems with delayed sampled outputs. Syst. Control Lett. 2016, 98, 86–96. [Google Scholar] [CrossRef]
  23. Shen, S.; Song, A.; Li, H.; Li, T. Cascade predictor for a class of mechanical systems under large uncertain measurement delays. Mech. Syst. Signal Process. 2022, 167, 108536. [Google Scholar] [CrossRef]
  24. Ammari, O.; Giri, F.; Krstic, M.; Benabdelhadi, A.; Chaoui, F.; El Majdoub, K. Adaptive Observer Design for Heat PDEs with Sensor Delay and Parameter Uncertainties. In IEEE Transactions on Automatic Control; IEEE: Piscataway, NJ, USA, 2024. [Google Scholar]
  25. Murguia, C.; Fey, R.H.B.; Nijmeijer, H. Immersion and invariance observers with time-delayed output measurements. Commun. Nonlinear Sci. Numer. Simul. 2016, 30, 227–235. [Google Scholar] [CrossRef]
  26. Della Santina, C.; Truby, R.L.; Rus, D. Data–driven disturbance observers for estimating external forces on soft robots. IEEE Robot. Autom. Lett. 2020, 5, 5717–5724. [Google Scholar] [CrossRef]
  27. Hu, J.; Xiong, R. Contact force estimation for robot manipulator using semiparametric model and disturbance Kalman filter. IEEE Trans. Ind. Electron. 2017, 65, 3365–3375. [Google Scholar] [CrossRef]
  28. Wahrburg, A.; Bös, J.; Listmann, K.D.; Dai, F.; Matthias, B.; Ding, H. Motor-current-based estimation of cartesian contact forces and torques for robotic manipulators and its application to force control. IEEE Trans. Autom. Sci. Eng. 2017, 15, 879–886. [Google Scholar] [CrossRef]
  29. Dehghan, S.A.M.; Koofigar, H.R.; Sadeghian, H.; Ekramian, M. Observer-based adaptive force–position control for nonlinear bilateral teleoperation with time delay. Control Eng. Pract. 2021, 107, 104679. [Google Scholar] [CrossRef]
  30. Guajardo-Benavides, E.J.; Arteaga, M.A. On the finite time force estimation for bilateral teleoperation of robot manipulators with time varying delays. Control Eng. Pract. 2023, 138, 105622. [Google Scholar] [CrossRef]
  31. Mohammadi, A.; Marquez, H.J.; Tavakoli, M. Nonlinear disturbance observers: Design and applications to Euler Lagrange systems. IEEE Control Syst. Mag. 2017, 37, 50–72. [Google Scholar] [CrossRef]
  32. Liu, L.; Yue, X.; Wen, H.; Hao, Z.; Zhu, Y. A globally exponentially stable observer for torque estimation with only position measurement. Proc. Inst. Mech. Eng. Part I J. Syst. Control. Eng. 2024, 238, 238–251. [Google Scholar] [CrossRef]
  33. Sadeghian, H.; Keshmiri, M.; Villani, L.; Siciliano, B. Null-space impedance control with disturbance observer. In Proceedings of the 2012 IEEE/RSJ International Conference on Intelligent Robots and Systems, Vilamoura-Algarve, Portugal, 7–12 October 2012; IEEE: Piscataway, NJ, USA, 2012; pp. 2795–2800. [Google Scholar]
  34. Sun, T.; Cheng, L.; Wang, W.; Pan, Y. Semiglobal exponential control of Euler–Lagrange systems using a sliding-mode disturbance observer. Automatica 2020, 112, 108677. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Edwards, C.; Belmont, M.; Li, G. Robust model predictive control for constrained linear system based on a sliding mode disturbance observer. Automatica 2023, 154, 111101. [Google Scholar] [CrossRef]
  36. Zhao, K.; Jia, N.; She, J.; Dai, W.; Zhou, R.; Liu, W.; Li, X. Robust model-free super-twisting sliding-mode control method based on extended sliding-mode disturbance observer for PMSM drive system. Control Eng. Pract. 2023, 139, 105657. [Google Scholar] [CrossRef]
  37. Moreno, J.A.; Osorio, M. A Lyapunov approach to second-order sliding mode controllers and observers. In Proceedings of the 2008 47th IEEE Conference on Decision and Control, Cancun, Mexico, 9–11 December 2008; IEEE: Piscataway, NJ, USA, 2008; pp. 2856–2861. [Google Scholar]
  38. Obuz, S.; Parikh, A.; Chakraborty, I.; Dixon, W. Lyapunov-based control of an uncertain Euler-Lagrange system with uncertain time-varying input delays without delay rate constraints. IFAC-PapersOnLine 2016, 49, 141–146. [Google Scholar] [CrossRef]
  39. Shen, S.; Fu, L.; Song, A.; Li, T. Sequential Predictors for Uncertain Euler–Lagrange Systems with Large Transmission Delays. Fractal Fract. 2023, 7, 244. [Google Scholar] [CrossRef]
  40. Kamalapurkar, R.; Fischer, N.; Obuz, S.; Dixon, W.E. Time-varying input and state delay compensation for uncertain nonlinear systems. IEEE Trans. Autom. Control 2015, 61, 834–839. [Google Scholar] [CrossRef]
  41. Hua, C.; Yang, Y. Bilateral Teleoperation Design with/without Gravity Measurement. IEEE Trans. Instrum. Meas. 2012, 61, 3136–3146. [Google Scholar] [CrossRef]
  42. Tutsoy, O.; Barkana, D.E. Model free adaptive control of the under-actuated robot manipulator with the chaotic dynamics. ISA Trans. 2021, 118, 106–115. [Google Scholar] [CrossRef]
Figure 1. Control block diagram of remote robotic system.
Figure 1. Control block diagram of remote robotic system.
Symmetry 16 00975 g001
Figure 2. Torque estimation results of two joints.
Figure 2. Torque estimation results of two joints.
Symmetry 16 00975 g002
Figure 3. State reconstruction results of joint angles.
Figure 3. State reconstruction results of joint angles.
Symmetry 16 00975 g003
Figure 4. State reconstruction results of joint velocities.
Figure 4. State reconstruction results of joint velocities.
Symmetry 16 00975 g004
Figure 5. Estimation errors e 2 of input torque in different situations.
Figure 5. Estimation errors e 2 of input torque in different situations.
Symmetry 16 00975 g005
Figure 6. Reconstruction errors ε 1 of joint angles in different situations.
Figure 6. Reconstruction errors ε 1 of joint angles in different situations.
Symmetry 16 00975 g006
Figure 7. Reconstruction errors ε 2 of joint velocities in different situations.
Figure 7. Reconstruction errors ε 2 of joint velocities in different situations.
Symmetry 16 00975 g007
Figure 8. Reconstruction errors ε 1 of joint angles under different time delays.
Figure 8. Reconstruction errors ε 1 of joint angles under different time delays.
Symmetry 16 00975 g008
Figure 9. Comparison results of reconstruction errors ε 1 by using different methods.
Figure 9. Comparison results of reconstruction errors ε 1 by using different methods.
Symmetry 16 00975 g009
Table 1. Integral square error of ε 1 compared with conventional high-gain predictor.
Table 1. Integral square error of ε 1 compared with conventional high-gain predictor.
MethodISEPercentage
Proposed0.265586%
Without torque estimation0.6250202%
High-gain predictor0.3099100%
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

Shen, S.; Jin, L. State Reconstruction of Remote Robotic System Using Delayed Output and Torque Estimation. Symmetry 2024, 16, 975. https://doi.org/10.3390/sym16080975

AMA Style

Shen S, Jin L. State Reconstruction of Remote Robotic System Using Delayed Output and Torque Estimation. Symmetry. 2024; 16(8):975. https://doi.org/10.3390/sym16080975

Chicago/Turabian Style

Shen, Shaobo, and Liqiang Jin. 2024. "State Reconstruction of Remote Robotic System Using Delayed Output and Torque Estimation" Symmetry 16, no. 8: 975. https://doi.org/10.3390/sym16080975

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