1. Introduction
Nonlinear bilateral teleoperation systems can expand the range of perception, and enable humans to complete complex tasks in a remote operating environment. A representative nonlinear bilateral teleoperation system consists of the following five parts: human operator, master robot, the network communication channel, slave robot, and remote environment. In recent years, the potential applications of teleoperation systems are in the various fields, i.e., the remote handling of hazardous materials [
1], underwater maintenance and repairing tasks [
2], space exploration [
3], telemedicine [
4], and so on. In all of these applications, the tasks occur in long-distance and/or hazardous environments. However, due to the limitation of bandwidth for the communication, there will be inevitable time delays. As we all know, time delay is one of the factors influencing the stability of nonlinear bilateral teleoperation systems. In practice, because of the extremely complicated nonlinearity of the nonlinear bilateral teleoperation systems and the certain constraint conditions on their outputs or states, nonlinear bilateral teleoperation systems havedifficultly in performing ideal performances.
Nowadays, a series of effective control strategies have been performed to solve the stability problems of nonlinear bilateral teleoperation systems. In [
5], a notion of wave variable had been developed to handle the time delay issue. In [
6], the instability, caused by time delays, had been conquered with a new passivity and scattering theory. In [
7], a robust adaptive control algorithm is developed to deal with the system’s uncertainties and to provide a smooth estimation of delayed reference signals. In [
8], Baranitha tackled the time-varying delay problem for a single-master multi-slave teleoperation system by assuming an asymmetric and semi-Markovian jump protocol for communication of the slaves with the master. There were also the passivity-based assumptions employed to ensure the stability of the position and velocity of the teleoperation systems, such as [
9,
10]. After that, Lu proposes a relative impedance-based force control method for a bimanual robot teleoperation system with varying time delays in [
11]. Lu adopts a strategies that design a hybrid error item to minimize both the position and force tracking errors. Additionally, in order to reduce the tracking error and ensure the stability of the system, an asymmetrical wave variable compensation method is proposed [
12], where the forward wave variable compensates the backward wave variable. However, the abovementioned research methods only work for joint space-based teleoperation systems. When the slave and the master are kinematically different; for instance, the slave robot is bigger/smaller than another one. Previous controllers in the joint space cannot bring about a satisfactory working performance. In order to make the robot end effector reach the desired position in the task space, Takegaki and Arimoto proposed a new position control method-based Jacobian task space transition matrix [
13]. With the development of this study, many methodologies have been widely described in the literature to aim at task space-based teleoperation system control such as the new nonlinear saturated proportional derivative (SPD) control strategy based on approximate Jacobian matrix [
14], a novel (nP+D)-like controller for task-space tracking performance [
15], nonlinear adaptive saturation control strategy with limited amplitude [
16], adaptive control schemes based on assumed parametric linearization of kinematics and dynamics [
17], and so on. It is worth noting that there is no full discussion about the system uncertainties and external disturbances in the above literature, although the system stability and synchronization performances were achieved. Indeed, the above-mentioned consequences are performed under certain assumptions where the models for the communication network and the master/slave controllers have been supposed to be in the continuous-time domain. In reality, the communication media are composed of undependable communication networks that may have variable delays, lost or reordered packets. Therefore, there is an urgent need for nonlinear bilateral teleoperation in developing a discrete-time theory.
It is well known that sliding mode control (SMC) has effective control performance for both linear and nonlinear systems, and provides more noticeable robustness and simpler calculation than other robust control approaches [
18]. However, the computation of the computer is based on discrete sample systems. If the continuous-time SMC algorithms are applied directly to discrete-time systems, it is of concern that there would be some indomitable problems such as chattering and discretization errors. Therefore, the design of discrete sliding mode control (DSMC) has attracted a lot of attention in recent years. There are also some contributions in the area of DSMC such as [
19]. Better performance could be guaranteed by considering a sampling period in the design phase, even if the sampling period was quite long [
20]. In [
21], Ma developed a novel dead zone sliding mode reaching law with disturbance compensation for uncertain discrete-time systems. A new adaptive second order DSMC scheme is proposed for a class of uncertain nonlinear systems [
22]; however, better robustness and trace performance come at the cost of slightly more complex control logic. In [
23], an optimal sliding surface coupled with a delay predictor had been presented to construct a DSMC for overcoming system parametric uncertainty. Moreover, we concentrate particularly on the active disturbance rejection control (ADRC). In the ADRC strategy, the extended state observer (ESO) is adopted to obtain the real-time estimation of the lumped system uncertainties including both internal unmodeled dynamics and external disturbance in the system [
24]. In [
25], a third-order super-twisting extended state observer was proposed, which enhanced the estimation convergence and provided strong disturbance estimation against fast speed and load variation. Liu [
26] proposed that an improved model predictive control (MPC) combined with extended state observe exhibits better control performance and faster dynamic response, where the ESO provides real-time disturbance compensation for the prediction control algorithm. Consequently, in this paper, we propose a reasonable method to effectively handle total system disturbances is by introducing ESOs for the uncertain teleoperation system.
Since the communication between the master and the slave is achieved through a packet-switched network, this paper focuses specifically on the synchronization control for discrete-time nonlinear bilateral teleoperation system in task space, which includes internal unmodeled dynamics, external disturbance and time-delays. Therefore, this paper is aimed at developing a new discrete sliding mode control algorithm to ensure the synchronization performance between the master and the slave via applying the adaptive extended state observer (AESO) to compensate total disturbances. The main accomplishments of the proposed strategy are summarized as follows: (i) with the aim of avoiding discretization after control design, a discrete sliding mode control algorithm is designed, in order to consider the discrete states caused by unreliable communication networks; (ii) a new reaching law of DSMC is developed to reduce the chattering while ensuring the tracking error quickly converges to zero domain; (iii) The parameter of AESO is designed so that the estimation errors quickly converge to smaller neighborhood and experimental comparisons demonstrate better. Finally, the proposed control method is simulated and tested by an experiment executed on a nonlinear bilateral teleoperation system composed of two Phantom Premium 1.5A robot manipulators. The test results reveal that the proposed control algorithm yields remarkable control performance.
The rest of this paper is organized as following. The concerned background is discussed in
Section 2. In
Section 3, we propose AESO-based DSMC and present the stability analyses of the nonlinear bilateral teleoperation system based on the designed controllers.
Section 4 shows the simulation and experiment results. The concluding remarks are given in
Section 5.
2. Problem Statement and Preliminaries
In this section, a
-degree-of-freedom (DOF) master-slave nonlinear bilateral telerobotic system is considered as follows:
where
and
denote the master side and the slave side of the nonlinear bilateral teleoperation system, respectively,
with
is the joint position vector,
is the joint velocity vector,
is the joint accelerated velocity vector,
is the positive-definite inertia matrix,
is the matrix of centripetal and Coriolis term,
is the gravitational vector,
is the viscous friction vector,
denotes the unknown bounded external disturbance,
is control torque and
,
is applied input control vector,
is Jacobian matrix, and
denote master operator force and external environmental force, respectively.
We review the properties [
27,
28,
29] for teleoperation system as follows:
Property 1. The inertia matrix is a symmetric positive-definite matrix, and there are two positive constants and such that .
Property 2. There exists a positive scalar such that , with regard to all .
Property 3. If and are bounded, the time derivative of the term is also bounded.
Assumption 1. In this article, the Jacobian matrix is supposed to be known and that the robot arms are working in a finite task space; in other words, the Jacobian matrix has full rank.
Let
represent the task coordinates of the task-space end effectors. The positional relationship between task space and joint space can be represented by the following relation
where
,
represents the mapping relationship from joint space to task space, and
.
Then,
are expressed as follows:
After the model transformation, in the task space, the system (1) can be rewritten as:
where
,
In most practical applications, the precise mode cannot be obtained directly, due to the noise, friction, viscous friction, uncertain disturbances and so on. Due to the existence of certain uncertainties in
and
, the system (4) is rewritten as follows:
where
,
,
,
.
,
represent the nominal parts, while
and
represent the uncertainties. When
,
, otherwise,
. In this paper,
denotes the lumped system uncertainty, and it is assumed to be bounded.
Introducing the state vector
,
, let
,
,
,
, then the system (5) is transformed as follows:
Furthermore, the system (6) is rewritten as follows:
where
,
,
,
.
Notice that all the terms proposed in (6) and (7) can be easily calculated from (5). With the sampling time
, the discretization of the uncertain model equation is given as:
where
3. Main Results
This section addresses a presentation of a proposed control scheme based on discrete-time nonlinear bilateral teleoperation system, where the new AESO is introduced to estimate and compensate the uncertainty. Meanwhile, the development of a discrete-time SMC algorithm is depicted, with a view to settle the synchronization problem of bilateral teleoperation system in task space.
3.1. Adaptive Extended State Observer
In this part, the adaptive extended state observer will be employed for the system (9). Firstly, for the convenience of further analysis, the following state measurement values are introduced:
where
is the output to be controlled, and
is the one to be measured, which includes the measurement noise vector
. Then, the following assumptions for the observer are given as [
30]:
Assumption 2. is a white random sequence andwhere is a known matrix. Assumption 3. where is the estimation error of the AESO, and is a known matrix. Assumption 4. where , , is the discretized error and is a known diagonal matrix. Therefore, the linear structure of AESO is designed as follows:
where
,
, and
is the gain of the discrete AESO, ensuring that
and
can be employed as the estimation of
and
, respectively. Then, the original values of AESO (13) are taken as:
Therefore, the gain
of the AESO is designed as follows:
where
And
,
,
Define the estimation error
as follows:
then, we will provide the following Theorem 1 showing the property of the estimation error
.
Theorem 1. If Assumptions 2–4 hold, and there exist the positive real numbers , , , , , , , and as following:so that the inequality is fulfilled: Then the estimation error will be uniformly bounded in the mean square, if the initial estimation error satisfies: Before proof of Theorem 1, the following lemma is first discussed.
Lemma 1 ([31]).In the case of Theorem 1, there is a real number such that meet the inequality: For with given by (15).
Proof. From (19), and
, we have:
since, we get:
Because of
, invert both sides of this inequality (25), multiply both sides by
and
, then we finally obtain:
Proof of Theorem 1. From (7) and (13), we have:
where:
and the last term of Equation (28) satisfies:
Define
, thus:
With the above knowable, the measurement noise vector
and the estimation error
are unrelated, so the following inequality is given as:
According to
Assumption 4, the last third and second terms of (31) have the following upper bound:
Note that the proper
can make the two sides of the inequality as close as possible. If and only if
, the equation of (32) can be achieved. In consequence, we get the following equation:
where:
Because of
, the Equation (33) and Equation (34) indicate that:
Since
and
are diagonal matrices, we set:
□
Remark 1. For the parameter , if it is chosen as a constant, each item of gain is iterated in the same way. Therefore, it is possible that the corresponding gain for the derivative of total disturbance estimation is large, which could easily lead to an overgrowth of the total disturbance estimation. Consequently, the gain of controller would be set sufficiently large to stabilize the closed-loop teleoperation system. Unfortunately, this will come with the actuator saturation problem. In order to solve this problem, the new parameter in the form of a diagonal matrix, is proposed in this paper. In other words, parameters are adjusted, respectively, according to the variation of different system states in this paper. Therefore, each state could reach its ideal one as quickly as possible, then the stability will be guaranteed.
Thus:
and
, we get:
Therefore,
is regarded as the upper bound of the covariance matrix of
, and it is also recommended to be minimized by proposing
. Design
When the partial derivative of
with respect to
is zero, that is:
We design Lyapunov function as follows:
From (28), we have:
and using Lemma 1 and Equation (43), the Equation (44) can be transformed as:
Since
is not related to
, we can get further:
For the second and third terms of the (46), apply the (19) and we will have:
For the fourth term of the (46), apply the (19) and we will have:
and for the fifth term of the (46), apply the (19) and we will have
Since both sides of (49) are scalars, we can trace the right side of (49) and at the same time will not change its value, that is
Finally, synthesize the inequalities (46)–(50) above and we can get:
where
. When
, using (19) and (43), we obtain:
From discussion above, we may conclude that the appropriate positive parameters , , , , , , , and may be utilized to guarantee the stability of the error system (143), and the estimation error converges to a smaller zero domain.
3.2. Discrete Sliding Mode Surface
In the above subsection, the lumped system uncertainty of the bilateral teleoperation system is estimated by the new AESO (13). In this part, its estimation is employed as the compensation for the uncertainty, and DSMC is provided by the following main theorem. For the nonlinear bilateral teleoperation system, tracking trajectories of master and slave robots are
and
, respectively, where
is the constant time delay.
represents the time delay of signal transmission from the master robot to the slave robot and
stands for the time delay of signal transmission from the slave robot to the master robot. So, the position synchronization errors between the master and the slave are defined as follows:
Next, a switching surface is constructed, and the dynamic equation of the ideal sliding mode is obtained by using the equivalent control method. Then, sufficient conditions for the asymptotic dynamic stability of the sliding mode are obtained by using Lyapunov stability theory and LMI (Linear Matrix Inequality) method.
Select the following switching function:
with
,
, which is assumed that
. Then
According to the discrete sliding mode theory, the equation
is obtained when the system states reach the switching surface. Therefore, the ideal sliding mode equivalent control law can be obtained from formulas (8) and (55) as follows:
Note 1. The equivalent control law shown in Equation (57) is only used to analyze the stability of the sliding mode dynamic equation, and it is only used as a tool here. The actual sliding mode control law will be designed below.
By substituting the equivalent control law (56) into the system (8), the sliding mode dynamic equation of the uncertain discrete time-delay system under ideal conditions can be obtained as follows:
The stability of the sliding mode dynamic equation (58) is further analyzed by using the LMI technique, and sufficient conditions are given to guarantee the asymptotic stability of the sliding mode dynamic equation.
Theorem 2. For uncertain discrete time-delay system (8), switch function (54) is selected. If there exist positive definite matrix and scalar , , and the following LMI inequalities (58)-(60) are satisfied, then the sliding mode dynamic system (57) is asymptotically stable:where , , . Assumption 5. and represent the nonlinear part of the teleoperation system and the disturbance, so the following condition is satisfied:where , , and are known constant matrices, and satisfies . Lemma 2 ([32]). If and are the real matrixes with the appropriate dimensions, and satisfies . Then, for any non-zero constant , the following inequality exists: Lemma 3 ([32]). If and are the real matrixes with the appropriate dimensions, for any non-zero matrix , then Lemma 4 ([33]). (Schur’s theorem) For the following LMI inequalitywhere , , so this is the same thing as , . Proof. In this proof, to facilitate the proof process, the hypothesizes and lemmas above are given. □
Let us take the Lyapunov function as
Along the state trajectory of the system (57), it can be obtained that:
where
For the first and second terms of the expression
, it can be seen from the
Lemma 3 that:
Similarly, for other terms of the expression
, we can get:
According to the
Lemma 3, the following inequality is further obtained:
In addition, it can be obtained from
Assumption 5 that
where
,
. So, here we have
where:
Therefore, when
,
is true (If
). According to the
Lemma 4, the inequality
can be equivalent to:
Furthermore, the inequality (73) is equivalent to
where
According to the
Lemma 2, we can get:
Then, the following inequalities (76) and (77) can ensure that the (75) is true
where
.
Similarly, the inequality (76) is equivalent to
where
Then the following inequalities (79) and (80) can guarantee the existence of the (78):
where
.
Therefore, according to the Lemma 4, if LMI inequalities (58)–(60) are true, then . Therefore, . According to Lyapunov stability theory, the sliding mode dynamic system (57) is asymptotically stable. The proof process is over.
3.3. Discrete Sliding Mode Controller
This is an example of an equation: The control objective of this paper is that the system trajectory tracking error from any initial state can arrive at a switching surface
and reach the origin along the sliding surface. So, define a switching belt encircling sliding surface as follows:
where
,
denotes the kth coordinate in task space,
,
. The
is the width of the switching belt. In order to reduce chattering and improve dynamic quality in the sliding stage, a new reaching law, which consists of the index term
and the convergence parameter
is designed by
where
,
, is a switching gain,
, and
Remark 2. In practical applications, the tanh function is advocated to replace the sign function in order to reduce chattering, and a mass of experimental data also verifies its effectiveness. However, no one has theoretically analyzed the feasibility for the function. This paper will make an attempt in this area. Theoretical analysis will be presented in Theorem 3. According to expression (83), a conclusion can be drawn that if the system tracking error is far away from the discrete sliding surface, tends to . Thus tends to , which is less than , and tends to , which is greater than . On the contrary, when the system tracking error is close to the discrete sliding surface, tends to 1. Thus tends to , and tends to . Obviously, the index term is positive, therefore the stability of the nonlinear bilateral teleoperation system will be unaffected.
With consideration of the system (8) and the new reaching law (82), the proposed DSMC controller for the nonlinear bilateral teleoperation system can be given as
The control block diagram of nonlinear bilateral teleoperation system with controller (86) is displayed in
Figure 1.
For the design of the discrete sliding mode controller, the following lemma and assumption will be utilized.
Lemma 5 ([34]). For,, where. Assumption 6. According to the Theory 1, the disturbanceestimation erroris bounded and converges to a very small range. Therefore, it is reasonable to assume that:where , , , , and is the positive constant and denotes the upper bound of . Theorem 3. For the discrete nonlinear bilateral teleoperation system (89) based on the controller (84), under the Assumption 6, if the following condition is maintained: (a). The system trajectory tracking error from any initial state will enter this switching belt of DSMC defined by (b). Once the system trajectory tracking errors enter this switching belt , they cannot escape from it.
In this paper, because the coupling relationship between the states of nonlinear bilateral teleoperation system is compensated as the uncertainty during the AESO design phase, with regard to , the controllers are independent on each other and then the stability analysis can be demonstrated in the same way. Hereinafter, the stability analysis is discussed for only one.
Proof. In this proof, two cases will be considered, namely, the positive and negative values of
. The equivalent form of
Theorem 3 is represented as follows:
where
,
.
(a). Case 1: If
, due to
, it can be obtained from (82) and (86) that:
According to the
Lemma 5, then
Therefore, the sequence
is strictly monotonously decreasing when
. It is concluded that there must exist a positive integer
so that the following inequality (91) holds:
when
, the system trajectory tracking error enters the switching belt
of DSMC.
Case 2: If
, due to
and the condition (86), it can be obtained that
Therefore, the sequence
is strictly monotonously increasing when
. It is concluded that there must exist a positive integer
so that the following inequality (93) holds:
when
, the system trajectory tracking error enters the switching belt
of DSMC.
Therefore, in view of (91) and (93), it can be concluded that if lies outside the switching belt defined by (87), then the system trajectory tracking error from any initial state will enter into this switching belt of DSMC.
(b). Case 1: When the
enters the switching belt
, namely
, then:
Suppose
has escaped the switching belt
again, namely,
, then
However, the result is inconsistent with the hypothesis
, so that the hypothesis is not set up. Hence,
Case 2: When the
enters the switching belt
defined by (87), namely,
, then
Suppose
has escaped the switching belt
again, namely,
, then
However, the result is inconsistent with the hypothesis
so that the hypothesis is not set up. Hence,
In short, , when . That is to say, once the system trajectory tracking error enters this switching belt , they cannot escape from it. □
4. Simulation and Experiment
In this section, the simulated analyses and experimental results on the strength of the proposed control strategy are presented, with the purpose of verifying the effectiveness of the discrete SMC algorithm and the superiority of its control performance. Firstly, the simulations are implemented on two three-degree-of-freedom robot arms
The kinematic relationship between task space and joint space of the nonlinear bilateral teleoperation system is given by
where
,
represent the link lengths and
represent the joint angles of the robot arms.
For the simulation, the relevant parameter values are set as
,
,
,
. The human force
is imposed on the master robot, which is depicted in
Figure 2. While in the slave site, the force is kept to zero. The simulation results are utilized to prove the following: (1) When the master robot moves, does the slave robot follow the master? (2) When the human-input force vanishes, does the tracking error between the master and slave vanish as well? The aim of (1) and (2) is to present the stability of the nonlinear bilateral teleoperation system.
Now, we evaluate the validity of the proposed control strategy for the nonlinear bilateral teleoperation system (100) with existence of both time delays and flexible friction forces. First, time delays are adjusted to
. The sampling period is
, and parameters of adaptive ESO are chosen as
Parameters of discrete-time sliding mode surface are set as , , , , , , .
Afterwards, the trajectory estimation errors of master and slave robots are exhibited in
Figure 3 and
Figure 4, respectively.
The disturbance estimation errors of the master and the slave are displayed in
Figure 5 and
Figure 6. As shown, estimation errors of both the motion trajectory and the disturbance converge to a very small range, and then approach rapidly to zero after
. Therefore, accurate estimation of total system disturbances is provided by AESO. In order to show the superiority of the proposed controller, PD controller (Proportional derivative controller) is taken as the comparison term. Now, the motion trajectories of the master and the slave end effectors with proposed DSMC are shown in
Figure 7. Obviously, the slave can accurately reproduce the trajectory of the master and the prominent synchronization performance is completed. On the contrary, in
Figure 8, at the beginning of the movement, the chattering is quite serious, and the position tracking effect of the master robot and the slave robot is poor.
To further prove the superiority of the controller designed in this paper, the control algorithm is implemented on the teleoperation experiment platform built in the laboratory. The nonlinear bilateral teleoperation system model includes a couple of Phantom Premium 1.5A robot arms (SensAble Technologies, Inc.) to be performed, which is shown in
Figure 9. In the experiment, the sampling period is
, and the parameters of the discrete sliding mode control algorithm are selected as
,
,
,
,
,
,
.
Then the parameters of adaptive ESO are chosen as
Before the experiment, the information transmission delay between the master and the slave robots by Simulink module is set to
. With the aim of verifying the performance superiority of the AESO designed in this paper, the contrast experiment between the proposed AESO and another compared one with a constant parameter
is provided. The experiment results are displayed in
Figure 10,
Figure 11,
Figure 12,
Figure 13,
Figure 14,
Figure 15 and
Figure 16. As shown in
Figure 10,
Figure 11,
Figure 12 and
Figure 13, the motion trajectory estimation errors and the total disturbance estimation of the master and the slave based on this paper and the compared one is shown. It is noticed that the trajectory estimation errors based on this paper converge within 0.02 m, while others based on the compared one converge within 0.3 m. Comparing the total disturbance estimation, the AESO designed in this paper could effectively estimate the total disturbance; whereas, another one based on the compared one brings a hundredfold increase in total disturbance estimation and thus loses the estimation performance.
In addition,
Figure 14 and
Figure 15 display the control torques of the master and the slave robots, respectively. The motion trajectory tracking between the master and the slave is shown in
Figure 16. We can see that the master robot stops moving after 50 s, and then the slave robot also promptly stops moving. Ultimately, the two robots stop in the same position.
Next, the communication time delay is increased to
ms. The trajectory estimations and the estimation errors for the master and the slave based on this paper are displayed in
Figure 17,
Figure 18,
Figure 19 and
Figure 20. The control torques of the master and the slave robots are shown in
Figure 21 and
Figure 22, respectively. Furthermore,
Figure 23 presents the motion trajectory tracking between the master and the slave robots. From this experiment, although the slave robot responds slowly due to the increased time delay, it still follows the master motion. The experimental results and analyses further validate the superiority and validity of the proposed control strategy in this paper.