Abstract

Robust force/motion control strategies are presented for mobile manipulators under both holonomic and nonholonomic constraints in the presence of uncertainties and disturbances. The controls are based on structural knowledge of the dynamics of the robot, and the actuator dynamics is also taken into account. The proposed control is robust not only to structured uncertainty such as mass variation but also to unstructured one such as disturbances. The system stability and the boundness of tracking errors are proved using Lyapunov stability theory. The proposed control strategies guarantee that the system motion converges to the desired manifold with prescribed performance. Simulation results validate that not only the states of the system asymptotically converge to the desired trajectory, but also the constraint force asymptotically converges to the desired force.

1. Introduction

Mobile manipulators refer to robotic manipulators mounted on mobile platforms. Such systems combine the advantages of mobile platforms and robotic arms and reduce their drawbacks [14]. For instance, the mobile platform extends the arm workspace, whereas the arm offers much operational functionality. Applications for such systems could be found in mining, construction, forestry, planetary exploration, teleoperation, and military [511].

Mobile manipulators possess complex and strongly coupled dynamics of mobile platforms and manipulators [1216]. A control approach by nonlinear feedback linearization was presented for the mobile platform so that the manipulator is always positioned at the preferred configurations measured by its manipulability [17]. In [14], the effect of the dynamic interaction on the tracking performance of a mobile manipulator was studied, and nonlinear feedback control for the mobile manipulator was developed to compensate the dynamic interaction. In [18], a basic framework for the coordination and control of vehicle-arm systems was presented, which consists of two basic task-oriented control: end-effector task control and platform self-posture control. The standard definition of manipulability was generalized to the case of mobile manipulators, and the optimization of criteria inherited from manipulability considerations were given to generate the controls of the system when its end-effector motion was imposed [19]. In [20], a unified model for mobile manipulator was derived, and nonlinear feedback was applied to linearize and decouple the model, and decoupled force/position control of the end-effector along the same direction for mobile manipulators was proposed and applied to nonholonomic cart pushing. The previously mentioned literature concerning with control of the mobile manipulator requires the precise information on the dynamics of the mobile manipulator; there may be some difficulty in implementing them on the real system in practical applications.

Different researchers have investigated adaptive controls to deal with dynamics uncertainty of mobile manipulators. Adaptive neural-network- (NN-) based controls for the arm and the base had been proposed for the motion control of a mobile manipulator [21, 22]; each NN control output comprises a linear control term and a compensation term for parameter uncertainty and disturbances. Adaptive control was proposed for trajectory/force control of mobile manipulators subjected to holonomic and nonholonomic constraints with unknown inertia parameters [23, 24], which ensures the state of the system to asymptotically converge to the desired trajectory and force. The principal limitation associated with these schemes is that controllers are designed at the velocity input level or torque input level, and the actuator dynamics are excluded.

As demonstrated in [2527], actuator dynamics constitute an important component of the complete robot dynamics, especially in the case of high-velocity movement and highly varying loads. Many control methods have therefore been developed to take into account the effects of actuator dynamics (see, e.g., [2830]). However, the literature is sparse on the control of the nonholonomic mobile manipulators including the actuator dynamics. In most of the research works for controlling mobile manipulators, joint torques are control inputs though in reality joints are driven by actuators (e.g., DC motors), and therefore using actuator input voltages as control inputs is more realistic. To this effect, actuator dynamics is combined with the mobile manipulator's dynamics in this paper.

This paper addresses the problem of stabilization of force/motion control for a class of mobile manipulator systems with both holonomic and nonholonomic constraints in the parameter uncertainties and external disturbances.

Unlike the force/motion control presented in [3137], which is proposed for the mechanical systems subject to either holonomic or nonholonomic constraints, in our paper, the control is to deal with the system subject to both holonomic and nonholonomic constraints. After the dynamics based on decoupling force/motion is first presented, the robust motion/force control is proposed for the system under the consideration of the actuator dynamics uncertainty to complete the trajectory/force tracking. The paper has main contributions listed as follows.(i)Decoupling robust motion/force control strategies are presented for mobile manipulator with both holonomic and nonholonomic constraints in the parameter uncertainties and external disturbances, and nonregressor-based control design is developed in a unified manner without imposing any restriction on the system dynamics.(ii)The actuators (e.g., DC motor) dynamics of both the mobile platform and the arm are integrated with mobile manipulator dynamics and kinematics so that the actuator input voltages are the control inputs thus making the system more realistic.

Simulation results are described in detail that show the effectiveness of the proposed control law.

The rest of the paper is organized as follows. The system description of mobile manipulator subject to nonholonomic constraints and holonomic is briefly described in Section 2. Problem statement for the system control is given in Section 4. The main results of robust adaptive control design are presented in Section 5. Simulation studies are presented by comparison between the proposed robust control with nonrobust control in Section 6. Concluding remarks are given in Section 7.

2. System Description

Consider an 𝑛 DOF mobile manipulator with nonholonomic mobile base. The constrained mechanical system can be described as 𝑀(𝑞)̈𝑞+𝐶(𝑞,̇𝑞)̇𝑞+𝐺(𝑞)+𝑑(𝑡)=𝐵(𝑞)𝜏+𝑓,(2.1) where 𝑞=[𝑞1,,𝑞𝑛]𝑇𝑅𝑛 denote the generalized coordinates; 𝑀(𝑞)𝑅𝑛×𝑛 is the symmetric bounded positive definite inertia matrix; 𝐶(̇𝑞,𝑞)̇𝑞𝑅𝑛 denotes the Centripetal and Coriolis torques; 𝐺(𝑞)𝑅𝑛 is the gravitational torque vector; 𝑑(𝑡) denotes the external disturbances; 𝜏𝑅𝑚 is the control inputs; 𝐵(𝑞)𝑅𝑛×𝑚 is a full rank input transformation matrix and is assumed to be known because it is a function of fixed geometry of the system; 𝑓𝑅𝑚 denotes the vector of constraint forces; 𝐽𝑅𝑛×𝑚 is Jacobian matrix; 𝜆=[𝜆𝑛,𝜆]𝑅𝑚 is Lagrange multipliers corresponding to the nonholonomic and holonomic constraints.

The generalized coordinates may be separated into two sets 𝑞=[𝑞𝑣,𝑞𝑎]𝑇, where 𝑞𝑣𝑅𝑣 describes the generalized coordinates for the mobile platform, 𝑞𝑎𝑅𝑟 is the coordinates of the manipulator, and 𝑛=𝑣+𝑟.

Assumption 2.1 (see [3840]). The mobile manipulator is subject to known nonholonomic constraints.

Assumption 2.2. The system (2.8) is subjected to 𝑘 independent holonomic constraints, which can be written as (𝑞)=0,(𝑞)𝑅𝑘,(2.2) where (𝑞) is full rank, then 𝐽(𝑞)=𝜕/𝜕𝑞.

Remark 2.3. In actual implementation, we can adopt the methods of producing enough friction between the wheels of the mobile platform and the ground such that this assumption holds [4143].

The vehicle is subjected to nonholonomic constraints, the 𝑙 nonintegrable and independent velocity constraints can be expressed as 𝐴𝑞𝑣̇𝑞𝑣=0,(2.3) where 𝐴(𝑞𝑣)=[𝐴𝑇1(𝑞𝑣),,𝐴𝑇𝑙(𝑞𝑣)]𝑇𝑅𝑣𝑅𝑙×𝑣 is the kinematic constraint matrix which is assumed to have full rank 𝑙. In the paper, the vehicle is assumed to be completely nonholonomic. The effect of the constraints can be viewed as a restriction of the dynamics on the manifold Ω𝑛 as Ω𝑛=𝑞𝑣,̇𝑞𝑣𝑞𝐴𝑣̇𝑞𝑣=0.(2.4)

The generalized constraint forces for the nonholonomic constraints can be given by 𝑓𝑛=𝐴𝑇𝑞𝑣𝜆𝑛.(2.5)

Assume that the annihilator of the codistribution spanned by the covector fields 𝐴1(𝑞𝑣),,𝐴𝑙(𝑞𝑣) is a (𝑣𝑙)-dimensional smooth nonsingular distribution Δ on 𝑅𝑣. This distribution Δ is spanned by a set of (𝑣𝑙) smooth and linearly independent vector fields 𝐻1(𝑞𝑣),,𝐻𝑣𝑙(𝑞𝑣); that is, Δ=span{𝐻1(𝑞𝑣),,𝐻𝑣𝑙(𝑞)}, which satisfy, in local coordinates, the following relation: 𝐻𝑇𝑞𝑣𝐴𝑇𝑞𝑣=0,(2.6) where 𝐻(𝑞𝑣)=[𝐻1(𝑞𝑣),,𝐻𝑛𝑣𝑙(𝑞𝑣)]𝑅𝑣×(𝑣𝑙). Note that 𝐻𝑇𝐻 is of full rank. Constraints (2.3) imply the existence of vector ̇𝜂𝑅𝑣𝑙 [44], such that ̇𝑞𝑣𝑞=𝐻𝑣̇𝜂.(2.7)

Considering the nonholonomic constraints (2.3) and its derivative, the dynamics of mobile manipulator can be expressed as 𝐻𝑇𝑀𝑣𝐻𝐻𝑇𝑀𝑣𝑎𝑀𝑎𝑣𝐻𝑀𝑎̈𝜂̈𝑞𝑎+𝐻𝑇𝑀𝑣̇𝐻+𝐻𝑇𝐶𝑣𝐻𝐻𝑇𝐶𝑣𝑎𝑀𝑎𝑣̇𝐻+𝐶𝑎𝑣𝐻𝐶𝑎̇𝜂̇𝑞𝑎+𝐻𝑇𝐺𝑣𝐺𝑎+𝐻𝑇𝑑𝑣𝑑𝑎=𝐻𝑇𝐵𝑣𝜏𝑣𝐵a𝜏𝑎+𝐽00𝑣𝐽𝑎𝑇0𝜆.(2.8)

From Assumption 2.2, the holonomic constraint force 𝑓 can be converted to the joint space as 𝑓=𝐽𝑇𝜆. Hence, the holonomic constraint on the robot’s end effector can be viewed as restricting only the dynamics on the constraint manifold Ω defined by Ω={(𝑞,̇𝑞)(𝑞)=0,𝐽(𝑞)̇𝑞=0}. The vector 𝑞𝑎 can be further rearranged and partitioned into 𝑞𝑎=[𝑞1𝑎,𝑞2𝑎]𝑇;  𝑞1𝑎𝑅𝑟𝑘 describes the constrained motion of the manipulator, and 𝑞2𝑎𝑅𝑘 denotes the remaining joint variable. Then, 𝐽(𝑞)=𝜕,𝜕𝜂𝜕𝜕𝑞1𝑎,𝜕𝜕𝑞2𝑎.(2.9) From [45], it could be concluded 𝑞 is the function of 𝜁=[𝜂,𝑞1𝑎]𝑇, that is, 𝑞=𝑞(𝜁), and we have ̇𝜁̇𝑞=𝐿(𝜁), where 𝐿(𝜁)=𝜕𝑞/𝜕𝜁, ̈̇̇𝜁̈𝑞=𝐿(𝜁)𝜁+𝐿(𝜁), and 𝐿(𝜁), 𝐽1(𝜁)=𝐽(𝑞(𝜁)) satisfy the relationship 𝐿𝑇(𝜁)𝐽1𝑇(𝜁)=0.(2.10) The dynamic model (2.8), when it restricted to the constraint surface, can be transformed into the reduced model: 𝑀1𝐿̈(𝜁)𝜁+𝐶1̇𝜁+𝐺1+𝑑1(𝑡)=𝑢+𝐽1𝑇𝜆,(2.11) where 𝑀1=𝐻𝑇𝑀𝑣𝐻𝐻𝑇𝑀𝑣𝑎𝑀𝑎𝑣𝐻𝑀𝑎,𝐶1=𝐻𝑇𝑀𝑣̇𝐻𝐻𝑇𝑀𝑣𝑎𝑀𝑎𝑣𝐻𝑀𝑎̇𝐻𝐿(𝜁)+𝑇𝑀𝑣̇𝐻+𝐻𝑇𝐶𝑣𝐻𝐻𝑇𝐶𝑣𝑎𝑀𝑎𝑣̇𝐻+𝐶𝑎𝑣𝐶𝑎𝐺𝐿(𝜁),1=𝐻𝑇𝐺𝑣𝐺𝑎,𝑑1𝐻(𝑡)=𝑇𝑑𝑣𝑑𝑎,𝑢=𝐵1𝜏,𝐵1=𝐻𝑇𝐵𝑣00𝐵𝑎𝜂𝑞,𝜁=1𝑎.(2.12)

Multiplying 𝐿𝑇 by both sides of (2.11), we can obtain 𝑀𝐿̈(𝜁)𝜁+𝐶𝐿̇𝜁̇𝜁,𝜁+𝐺𝐿+𝑑𝐿(𝑡)=𝐿𝑇𝐵1𝜏.(2.13) The force multipliers 𝜆 can be obtained by (2.11): 𝜆𝐶=𝑍(𝜁)1̇𝜁̇𝜁,𝜁+𝐺1+𝑑1(𝑡)𝐵1𝜏,(2.14) where 𝑀𝐿=𝐿𝑇𝑀1𝐿, 𝐶𝐿=𝐿𝑇𝐶1, 𝐺𝐿=𝐿𝑇𝐺1, 𝑍=(𝐽1(𝑀1)1𝐽1𝑇)1𝐽1(𝑀1)1.

Property 1. The matrix 𝑀𝐿 is symmetric and positive definite.

Property 2. The matrix ̇𝑀𝐿2𝐶𝐿 is skew symmetric.

Property 3 (see [46]). For holonomic systems, matrices 𝐽1(𝜁), 𝐿(𝜁) are uniformly bounded and uniformly continuous if 𝜁 is uniformly bounded and continuous, respectively.

Property 4. There exist some finite positive constants 𝑐𝑖>0(1𝑖4) and finite nonnegative constant 𝑐𝑖0(𝑖=5) such that for all 𝜁𝑅𝑛, for all ̇𝜁𝑅𝑛, 𝑀𝐿(𝜁)𝑐1, 𝐶𝐿̇(𝜁,𝜁)𝑐2+𝑐3̇𝜁, 𝐺𝐿(𝜁)𝑐4, and sup𝑡0𝑑𝐿(𝑡)𝑐5.

3. Actuator Dynamics

The joints of the mobile manipulators are assumed to be driven by DC motors. Consider the following notations used to model a DC motor: 𝜈𝑅𝑚 represents the control input voltage vector; 𝐼 denotes an 𝑚-element vector of motor armature current; 𝐾𝑁𝑅𝑚×𝑚 is a positive definite diagonal matrix which characterizes the electromechanical conversion between current and torque; 𝐿𝑎=diag[𝐿𝑎1,𝐿𝑎2,𝐿𝑎3,,𝐿𝑎𝑚], 𝑅𝑎=diag[𝑅𝑎1,𝑅𝑎2,𝑅𝑎3,,𝑅𝑎𝑚], 𝐾𝑒=diag[𝐾𝑒1,𝐾𝑒2,𝐾𝑒3,,𝐾𝑒𝑚], 𝜔=[𝜔1,𝜔2,,𝜔𝑚]𝑇 represent the equivalent armature inductances, resistances, back EMF constants, angular velocities of the driving motors, respectively; 𝐺𝑟=diag(𝑔𝑟𝑖)𝑅𝑚×𝑚 denotes the gear ratio for 𝑚 joints; 𝜏𝑚 are the torque exerted by the motor. In order to apply the DC servomotors for actuating an 𝑛-DOF mobile manipulator, assuming no energy losses, a relationship between the 𝑖th joint velocity ̇𝑞𝑖 and the motor shaft velocity 𝜔𝑖 can be presented as 𝑔𝑟𝑖=𝜔𝑖/̇𝑞𝑖=𝜏𝑖/𝜏𝑚𝑖 with the gear ratio of the 𝑖th joint 𝑔𝑟𝑖, the 𝑖th motor shaft torque 𝜏𝑚𝑖, and the 𝑖th joint torque 𝜏𝑖. The motor shaft torque is proportional to the motor current 𝜏𝑚=𝐾𝑁𝐼. The back EMF is proportional to the angular velocity of the motor shaft; then we can obtain 𝐿𝑎𝑑𝐼𝑑𝑡+𝑅𝑎𝐼+𝐾𝑒𝜔=𝑣.(3.1) In the actuator dynamics (3.1), the relationship between 𝜔 and ̇𝜁 is dependent on the type of mechanical system and can be generally expressed as 𝜔=𝐺𝑟𝑇̇𝜁.(3.2) The structure of 𝑇 depends on the mechanical systems to be controlled. For instance, in the simulation example, a two-wheel differential drive 2-DOF mobile manipulator is used to illustrate the control design. From [47], we have 𝑟̇𝜃𝑣=𝑙̇𝜃+𝑟𝑟2,̇𝑟̇𝜃𝜃=𝑟̇𝜃𝑟𝑙,̇𝜃2𝑙1=̇𝜃1,̇𝜃2=̇𝜃2,(3.3) where ̇𝜃𝑙 and ̇𝜃𝑟 are the angular velocities of the two wheels, respectively, and 𝑣 is the linear velocity of the mobile platform, as shown in Figure 1. Since ̇𝑦=𝑣cos𝜃, we have ̇𝜃𝑙̇𝜃𝑟̇𝜃1̇𝜃2𝑇̇𝜃̇𝜃=𝑇̇𝑦1̇𝜃2𝑇,1𝑇=𝑙𝑟cos𝜃𝑟100𝑙𝑟cos𝜃𝑟,0000100001(3.4) where 𝑟 and 𝑙 are shown in Figure 1.

Eliminating 𝜔 from the actuator dynamics (3.1) by substituting (3.2), one obtains 𝐿𝑇𝐵1𝐺𝑟𝐾𝑁𝐼=𝑀𝐿̈(𝜁)𝜁+𝐶𝐿̇𝜁̇𝜁,𝜁+𝐺𝐿+𝑑𝐿𝜆(𝑡),(3.5)𝐶=𝑍(𝜁)2̇𝜁+𝐺2+𝑑2(𝑡)𝐵1𝐺𝑟𝐾𝑁𝐼,(3.6)𝜈=𝐿𝑎𝑑𝐼𝑑𝑡+𝑅𝑎𝐼+𝐾𝑒𝐺𝑟𝑇̇𝜁.(3.7)

Until now we have brought the kinematics (2.3), dynamics (3.5), (3.6) and actuator dynamics (3.7) of the considered nonholonomic system from the generalized coordinate system 𝑞𝑅𝑛 to feasible independent generalized velocities 𝜁𝑅𝑛𝑙𝑘 without violating the nonholonomic constraint (2.3).

4. Problem Statement

Since the system is subjected to the nonholonomic constraint (2.3) and holonomic constraint (2.2), the states 𝑞𝑣, 𝑞1𝑎, 𝑞2𝑎 are not independent. By a proper partition of 𝑞𝑎, 𝑞2𝑎 is uniquely determined by 𝜁=[𝜂,𝑞1𝑎]𝑇. Therefore, it is not necessary to consider the control of 𝑞2𝑎.

Given a desired motion trajectory 𝜁𝑑(𝑡)=[𝜂𝑑𝑞1𝑎𝑑]𝑇 and a desired constraint force 𝑓𝑑(𝑡), or, equivalently, a desired multiplier 𝜆(𝑡), the trajectory and force tracking control is to determine a control law such that for any ̇(𝜁(0),𝜁(0))Ω, 𝜁, ̇𝜁, 𝜆 asymptotically converge to a manifold Ω𝑑 specified as Ω where Ω𝑑=̇𝜁,𝜁,𝜆𝜁=𝜁𝑑,̇̇𝜁𝜁=𝑑,𝜆=𝜆𝑑.(4.1)

The controller design will consist of two stages: (i) a virtual adaptive control input 𝐼𝑑 is designed so that the subsystems (3.5) and (3.6) converge to the desired values, and (ii) the actual control input 𝜈 is designed in such a way that 𝐼𝐼𝑑. In turn, this allows 𝜁𝜁𝑑 and 𝜆𝜆𝑑 to be stabilized to the origin.

Assumption 4.1. The desired reference trajectory 𝜁𝑑(𝑡) is assumed to be bounded and uniformly continuous and has bounded and uniformly continuous derivatives up to the second order. The desired Lagrangian multiplier 𝜆𝑑(𝑡) is also bounded and uniformly continuous.

5. Robust Control Design

5.1. Kinematic and Dynamic Subsystems

Let 𝑒𝜁=𝜁𝜁𝑑,  ̇𝜁𝑟=̇𝜁𝑑𝑘𝜁𝑒𝜁,  𝑟=̇𝑒𝜁+𝑘𝜁𝑒𝜁 with 𝑘𝜁>0,  𝑒𝛽=𝜆𝜆𝑑. A decoupled control scheme is introduced to control generalized position and constraint force separatively.

Consider the virtual control input 𝐼 is designed as 𝐼=𝐾𝑁1𝐺𝑟1𝐵11𝜏.(5.1) Let the control 𝑢 be as the form 𝑢=𝐿+𝑇𝑢𝑎𝐽1𝑇𝑢𝑏,𝑢𝑎=𝐵1𝐺𝑟𝐾𝑁𝑎𝐼𝑎,𝑢𝑏=𝐵1𝐺𝑟𝐾𝑁𝑏𝐼𝑏,(5.2) where 𝑢𝑎, 𝐼𝑎𝑅𝑛𝑙𝑘 and 𝑢𝑏, 𝐼𝑏𝑅𝑘 and 𝐿+𝑇=(𝐿𝑇𝐿)1𝐿𝑇. Then, (2.13) and (2.14) can be changed to 𝑀𝐿̈(𝜁)𝜁+𝐶𝐿̇𝜁̇𝜁,𝜁+𝐺𝐿+𝑑𝐿(𝑡)=𝐵1𝐺𝑟𝐾𝑁𝑎𝐼𝑎,𝐶(5.3)𝑍(𝜁)1̇𝜁̇𝜁,𝜁+𝐺1+𝑑1(𝑡)𝐿+𝑇B1𝐺𝑟𝐾𝑁𝑎𝐼𝑎+𝐵1𝐺𝑟𝐾𝑁𝑏𝐼𝑏=𝜆.(5.4) Consider the following control laws: 𝐵1𝐺𝑟𝐾𝑁𝑎𝐼𝑑𝑎=𝐾𝑝𝑟𝐾𝑖𝑟𝑑𝑡𝑟Φ2,Φ𝛾(𝑟)+𝛿(5.5)Φ=𝐶𝑇𝐵Ψ,(5.6)1𝐺𝑟𝐾𝑁𝑏𝐼𝑑𝑏=𝜒2𝜒+𝛿+𝜆𝑑𝐾𝑓𝑒𝜆,(5.7)𝜒=𝑐1𝐿𝑍(𝜁)+𝑇𝑑̇𝜁𝑑𝑡𝑑,(5.8) where 𝑐𝐶=1𝑐2𝑐3𝑐4𝑐5; ̇𝜁Ψ=(𝑑/𝑑𝑡)[𝑟̇𝜁]𝑟̇̇𝜁𝜁𝑟11𝑇; 𝐾𝑝,𝐾𝑖,𝐾𝑓 are positive definite. 𝛾(𝑟) can be defined as follows: if 𝑟𝜌, 𝛾(𝑟)=𝜌, else 𝛾(𝑟)=𝑟, 𝜌 is a small value, 𝛿(𝑡) is a time-varying positive function converging to zero as 𝑡, such that 𝑡0𝛿(𝜔)𝑑𝜔=𝑎<. There are many choices for 𝛿(𝑡) that satisfies the condition.

5.2. Control Design at the Actuator Level

Till now, we have designed a virtual controller 𝐼 and 𝜁 for kinematic and dynamic subsystems. 𝜁 tending to 𝜁𝑑 can be guaranteed, if the actual input control signal of the dynamic system 𝐼 be of the form 𝐼𝑑 which can be realized from the actuator dynamics by the design of the actual control input 𝜈. On the basis of the above statements we can conclude that if 𝜈 is designed in such a way that 𝐼 tends to 𝐼𝑑, then (𝜁𝜁𝑑)0 and (𝜆𝜆𝑑)0.

Defining 𝐼=𝑒𝐼+𝐼𝑑 and substituting 𝐼 and ̇𝜁 of (3.7) one gets 𝐿𝑎̇𝑒𝐼+𝑅𝑎𝑒𝐼+𝐾𝑒𝐺𝑟𝑇̇𝑒𝜁=𝐿𝑎̇𝐼𝑑𝑅𝑎𝐼𝑑𝐾𝑒𝐺𝑟𝑇̇𝜁𝑑+𝜈.(5.9)

The actuator parameters 𝐾𝑁, 𝐿𝑎, 𝑅𝑎, and 𝐾𝑒 are considered unknown for control design; however, there exist 𝐿0, 𝑅0, and 𝐾𝑒0, such that 𝐿𝑎𝐿0𝛼1,𝑅𝑎𝑅0𝛼2,𝐾𝑒𝐾𝑒0𝛼3.(5.10)

Consider the robust control law 𝜈=𝜈03𝑖=1𝑒𝐼𝜇2𝑖𝑒𝐼𝜇𝑖+𝛿𝐾𝑑𝑒𝐼,(5.11) where 𝜈0=𝐿0̇𝐼𝑑+𝑅0𝐼𝑑+𝐾𝑒0𝐺𝑟𝑇̇𝜁𝑑,𝜇1=𝛼1𝑑𝐼𝑑𝑡𝑑,𝜇2=𝛼2𝐼𝑑,𝜇3=𝛼3𝑑𝜁𝑑𝑡𝑑.(5.12)

5.3. Stability Analysis for the System

Theorem 5.1. Consider the mechanical system described by (2.1), (2.3), and (2.2); using the control law (5.5) and (5.7), the following hold for any (𝑞(0),̇𝑞(0))Ω𝑛Ω:(i)𝑟 and 𝑒𝐼 converge to a set containing the origin with the convergence rate as 𝑡; (ii)𝑒𝑞 and ̇𝑒𝑞 asymptotically converge to 0 as 𝑡;(iii)𝑒𝜆 and 𝜏 are bounded for all 𝑡0.

Proof. (i) By combing (3.5) with (5.5), the closed-loop system dynamics can be rewritten as 𝑀𝐿̇𝑟=𝐵1𝐺𝑟𝐾𝑁𝑎𝐼𝑑𝑎+𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼𝑀𝐿̈𝜁𝑟+𝐶𝐿̇𝜁𝑟+𝐺𝐿+𝑑𝐿𝐶𝐿𝑟.(5.13)
Substituting (5.5) into (5.13), the closed-loop dynamic equation is obtained: 𝑀𝐿̇𝑟=𝐾𝑝𝑟𝐾𝑖𝑟𝑑𝑡𝑟Φ2Φ𝛾(𝑟)+𝛿𝜇𝐶𝐿𝑟+𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼,(5.14) where 𝜇=𝑀𝐿̈𝜁𝑟+𝐶𝐿̇𝜁𝑟+𝐺𝐿+𝑑𝐿.
Consider the function 𝑉=𝑉1+𝑉2,𝑉1=12𝑟𝑇𝑀𝐿1𝑟+2𝑟𝑑𝑡𝑇𝐾𝑖𝑟𝑑𝑡+𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝𝑒𝜁,𝑉2=12𝑒𝑇𝐼𝐾𝑁𝑎𝐿𝑎𝑒𝐼.(5.15) Then, differentiating 𝑉1 with respect to time, we have ̇𝑉1=𝑟𝑇𝑀𝐿1̇𝑟+2̇𝑀𝐿𝑟+𝐾𝑖𝑟𝑑𝑡+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝̇𝑒𝜁.(5.16) From Property 1, we have (1/2)𝜆min(𝑀𝐿)𝑟𝑇𝑟𝑉(1/2)𝜆max(𝑀𝐿)𝑟𝑇𝑟. By using Property 2, the time derivative of 𝑉 along the trajectory of (5.14) is ̇𝑉1=𝑟𝑇𝐾𝑝𝑟𝑟𝑇𝜇𝑟2Φ2Φ𝛾(𝑟)+𝛿+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝̇𝑒𝜁+𝑟𝑇𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼𝑟𝑇𝐾𝑝𝑟𝑟2Φ2+Φ𝛾(𝑟)+𝛿𝑟Φ+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝̇𝑒𝜁+𝑟𝑇𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼𝑟𝑇𝐾𝑝𝑟𝑟2Φ2𝛾(𝑟)Φ2𝑟𝑟Φ𝛿Φ𝛾(𝑟)+𝛿+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝̇𝑒𝜁+𝑟𝑇𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼,(5.17) when 𝑟𝜌; therefore, ̇𝑉1𝑟𝑇𝐾𝑝𝑟+𝛿+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑑𝑟2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝𝑘𝜁𝑒𝜁+𝑟𝑇𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼.(5.18) Differentiating 𝑉2(𝑡) with respect to time, using (3.7), one has ̇𝑉2=𝑒𝑇𝐼𝐾𝑁𝑎𝐿𝑎̇𝐼𝑑𝑎+𝑅𝑎𝐼𝑑𝑎+𝐾𝑒𝐺𝑟𝑇̇𝜁𝑑+𝑅𝑎𝑒𝐼+𝐾𝑒𝐺𝑟𝑇̇𝑒𝜁𝜈.(5.19) Substituting 𝜈 in (5.19) by the control law (5.11), one has ̇𝑉2=𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑑+𝑅𝑎𝑒𝐼𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇̇𝑒𝜁𝑒𝑇𝐼𝐾𝑁𝑎𝐿𝑎𝐿0̇𝐼𝑑𝑒𝑇𝐼𝐾𝑁𝑎𝑅𝑎𝑅0𝐼𝑑𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐾𝑒0𝐺𝑟𝑇̇𝜁𝑑𝑒𝑇𝐼𝐾3𝑁𝑎𝑖=1𝜇2𝑖𝑒𝐼𝑒𝐼𝜇𝑖+𝛿𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑑+𝑅𝑎𝑒𝐼𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇̇𝑒𝜁+𝛼1𝐾𝑁𝑎𝑒𝐼̇𝐼𝑑+𝛼2𝐾𝑁𝑎𝑒𝐼𝐼𝑑𝑎+𝛼3𝐾𝑁𝑎𝐺𝑟𝑇𝑒𝐼𝜁𝑑𝐾3𝑁𝑎𝑖=1𝑒𝐼2𝜇2𝑖𝑒𝐼𝜇𝑖+𝛿𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑑+𝑅𝑎𝑒𝐼e𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇̇𝑒𝜁+𝐾3𝑁𝑎𝑖=1𝛼𝑖𝛿=𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑑+𝑅𝑎𝑒𝐼𝑒𝑇𝐼𝐾𝑁𝐾𝑒𝐺𝑟𝑇𝑟+𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇𝑘𝜁𝑒𝜁+𝐾𝑁𝑎𝛿3𝑖=1𝛼𝑖.(5.20)
Integrating (5.18) and (5.20), ̇𝑉 can be expressed as ̇𝑉𝑟𝑇𝐾𝑝𝑟+𝛿+2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝𝑟2𝑒𝑇𝜁𝑘𝜁𝐾𝑁𝑎𝐾𝑝𝑘𝜁𝑒𝜁+𝑟𝑇𝐵1𝐺𝑟𝐾𝑁𝑎𝑒𝐼𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑑+𝑅𝑎𝑒𝐼𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇𝑟+𝑒𝑇𝐼𝐾𝑁𝑎𝐾𝑒𝐺𝑟𝑇𝑘𝜁𝑒𝜁+𝐾𝑁𝑎𝛿3𝑖=1𝛼𝑖.(5.21)
We can obtain ̇𝑟𝑉𝑇𝑒𝜁𝑒𝐼𝑄𝐾𝑁𝑎000𝐾𝑁𝑎000𝐾𝑁𝑎𝑟𝑒𝜁𝑒𝐼,(5.22) where 𝐾𝑄=𝑝𝐾𝑝𝑘𝜁12𝐺𝑟𝐾𝑒𝑇𝐵1𝑘𝜁𝐾𝑝2𝑘𝜁𝐾𝑝𝑇𝑘𝜁12𝐾𝑒𝐺𝑟𝑇𝑘𝜁12𝐺𝑟𝐾𝑒𝑇𝐵112𝐾𝑒𝐺𝑟𝑇𝑘𝜁𝐾𝑑+𝑅𝑎.(5.23) The term 𝑄 on the right-hand side (5.22) can always be negative definite by choosing suitable 𝐾𝑝 and 𝐾𝑑. Since [𝐾𝑛𝑎] is positive definite, we only need to choose 𝐾𝑝 and 𝐾𝑑 such that 𝑄 is positive definite. Therefore, 𝐾𝑑 and 𝐾𝑝 can always be chosen to satisfy 𝐾𝑑+𝑅>𝐾𝑝112𝐺𝑟𝐾𝑒𝑇𝐵112𝐾𝑒𝐺𝑟𝑇𝑘𝜁2𝐼𝑘𝜁1𝑘𝜁1𝑘𝜁1𝑇1𝑘𝜁112𝐺𝑟𝐾𝑒𝑇𝐵112𝐾𝑒𝐺𝑟𝑇𝑘𝜁.(5.24)
If 𝑟𝜌, it is easy to obtain ̇𝑉0. 𝑟, 𝑒𝜁, and 𝑒𝐼 converge to a set containing the origin with 𝑡.
(ii) 𝑉 is bounded, which implies that 𝑟𝐿𝑛𝑘. From 𝑟=̇𝑒𝜁+𝑘𝜁𝑒𝜁, it can be obtained that 𝑒𝜁,̇𝑒𝜁𝐿𝑛𝑘. As we have established 𝑒𝜁,̇𝑒𝜁𝐿, from Assumption 4.1, we conclude that ̇̇𝜁𝜁(𝑡),𝜁(𝑡),𝑟̈𝜁(𝑡),𝑟(𝑡)𝐿𝑛𝑘 and ̇𝑞𝐿𝑛.
Therefore, all the signals on the right hand side of (5.14) are bounded, and we can conclude that ̇𝑟 and therefore ̈𝜁 are bounded. Thus, 𝑟0 as 𝑡 can be obtained. Consequently, we have 𝑒𝜁0, ̇𝑒𝜁0 as 𝑡. It follows that 𝑒𝑞,̇𝑒𝑞0 as 𝑡.
(iii) Substituting the control (5.5) and (5.7) into the reduced order dynamic system model (5.4) yields 1+𝐾𝑓𝑒𝜆𝐶=𝑍(𝜁)1̇𝜁̇𝜁,𝜁+𝐺1+𝑑1(𝑡)𝐿+𝑇𝐺𝑟𝐾𝑁𝑎𝐼𝑎+𝐵1𝐺𝑟𝐾𝑁𝑏𝐼𝑑𝑏+𝐵1𝐺𝑟𝐾𝑁𝑏𝑒𝐼=𝑍(𝜁)𝐿+𝑇𝑀𝐿̈𝜁+𝜒(𝜁)2𝜒+𝛿+𝐵1𝐺𝑟𝐾𝑁𝑏𝑒𝐼.(5.25) Since ̇𝜁=0 when 𝐼𝑅𝑘, (3.7) could be changed as 𝐿𝑎𝑑𝐼𝑏𝑑𝑡+𝑅𝑎𝐼𝑏=𝜈𝑏.(5.26) Therefore, 𝑟=0 and 𝑒𝜁=0 in the force space; (5.20) could be changed as ̇𝑉2=𝑒𝑇𝐼𝐾𝑁𝑏𝐾𝑑𝑒+𝑅𝐼+𝐾𝑁𝑏𝛿3𝑖=1𝛼𝑖.(5.27) Since 𝐾𝑁𝑏 is bounded, ̇𝑉<0, we can obtain 𝑒𝐼0 as 𝑡. The proof is completed by noticing that ̈𝜁, 𝑍(𝑞), 𝐾𝑁𝑏 and 𝑒𝐼 are bounded. Moreover, 𝜁𝜁𝑑, and 𝑍(𝜁)𝐿+𝑇𝑀𝐿̈𝜁(𝜁)(𝑑)+𝜒2/(𝜒+𝛿)𝛿, 𝑒𝐼0, the right-hand side terms of (5.25), tend uniformly asymptotically to zero; then it follows that 𝑒𝜆0, then 𝑓(𝑡)𝑓𝑑(𝑡).
Since 𝑟, 𝜁, ̇𝜁, 𝜁𝑟, ̇𝜁𝑟, ̈𝜁𝑟, 𝑒𝜆 and 𝑒𝐼 are all bounded, it is easy to conclude that 𝜏 is bounded from (5.2).

6. Simulations

To verify the effectiveness of the proposed control algorithm, let us consider a 2-DOF manipulator mounted on two-wheels-driven mobile base [23] shown in Figure 1. The mobile manipulator is subjected to the following constraints: ̇𝑥cos𝜃+̇𝑦sin𝜃=0. Using Lagrangian approach, we can obtain the standard form with 𝑞𝑣=[𝑥,𝑦,𝜃]𝑇, 𝑞𝑎=[𝜃1,𝜃2]𝑇, 𝑞=[𝑞𝑣,𝑞𝑎]𝑇, and 𝐴𝑣=[cos𝜃,sin𝜃,0]𝑇: 𝑀𝑣=𝑚𝑝12+2𝐼𝑤sin2𝜃𝑟22𝐼𝑤𝑟2sin𝜃cos𝜃𝑚12𝑑sin𝜃2𝐼𝑤𝑟2sin𝜃cos𝜃𝑚𝑝12+2𝐼𝑤cos2𝜃𝑟2𝑚12𝑑cos𝜃𝑚12𝑑sin𝜃𝑚12𝑑cos𝜃𝑀111,𝑀111=𝐼𝑝+𝐼12+𝑚12𝑑2+2𝐼𝑤𝐿2𝑟2,𝑀𝑎𝐼=diag12,𝐼2,𝑀𝑣𝑎=𝐼0.00.00.00.012,0.0𝐵=sin𝜃𝑟sin𝜃𝑟0.00.0cos𝜃𝑟cos𝜃𝑟𝑙0.00.0𝑟𝑙𝑟,𝐶0.00.00.00.01.00.00.00.00.01.0𝑣=2𝐼𝑤𝑟2̇𝜃sin𝜃cos𝜃2𝐼𝑤𝑟2̇𝜃sin2𝜃𝑚12𝑑̇𝜃cos𝜃0.02𝐼𝑤𝑟2̇𝜃cos2𝜃2𝐼𝑤𝑟2̇𝜃sin𝜃cos𝜃𝑚12𝑑̇,𝐶𝜃cos𝜃0.00.00.00.00.0𝑣𝑎=0.0,𝐶𝑎=0.0,𝐺𝑣=[]0.0,0.0,0.0𝑇,𝐺𝑎=0.0,𝑚2𝑔𝑙2sin𝜃2𝑇,,𝜏𝐻=tan𝜃0.00.00.01.00.00.00.00.01.00.00.00.00.01.00.00.00.00.01.0𝑣=𝜏𝑙,𝜏𝑟𝑇,𝜏𝑎=𝜏1,𝜏2𝑇,𝑚𝑝12=𝑚𝑝+𝑚12,𝑚12=𝑚1+𝑚2,𝐼12=𝐼1+𝐼2.(6.1)

Let the desired trajectory 𝑞𝑑=[𝑥𝑑,𝑦𝑑,𝜃𝑑,𝜃1𝑑,𝜃2𝑑]𝑇 and the end effector be subject to the geometric constraint Φ=𝑙1+𝑙2sin(𝜃2)=0, and 𝑦𝑑=1.5sin(𝑡), 𝜃𝑑=1.0sin(𝑡), 𝜃1𝑑=𝜋/4(1cos(𝑡)), 𝜆𝑑=10.0𝑁.

The trajectory and force tracking control problem is to design control law 𝜏 such that (4.1) holds and all internal signals are bounded.

In the simulation, we assume the parameter 𝑚𝑝=𝑚1=𝑚2=1.0, 𝐼𝑤=𝐼𝑝=1.0, 2𝐼1=𝐼2=1.0, 𝐼=0.5, 𝑑=𝐿=𝑅=1.0, 2𝑙1=1.0, 2𝑙2=0.6, 𝑞(0)=[0,2.0,0.6,0.5]𝑇, ̇𝑞(0)=[0.0,0.0,0.0,0.0]𝑇, 𝐾𝑁=diag[0.01], 𝐺𝑟=diag[100], 𝐿𝑎=[0.005,0.005,0.005,0.005]𝑇, 𝑅𝑎=[2.5,2.5,2.5,2.5]𝑇, and 𝐾𝑒=[0.02,0.02,0.02,0.02]𝑇. The disturbance on the mobile base is set 0.1sin(𝑡) and 0.1cos(𝑡). By Theorem 5.1, the control gains are selected as 𝐾𝑝=diag[1.0,1.0,1.0], 𝑘𝜁=diag[1.0,1.0,1.0], 𝐾𝑖=0.0 and 𝐾𝑓=0.995, 𝐶=[8.0,8.0,8.0,8.0,8.0]𝑇, 𝐾𝑁=0.1, 𝐾𝑑=diag[10,10,10,10], 𝛼1=0.008, 𝛼2=4.0, 𝛼3=0.03. The disturbance on the mobile base is set 0.1sin(𝑡) and 0.1cos(𝑡). The simulation results for motion/force are shown in Figures 2, 3, 4, 5, 6, 7, 8, and 9. The desired currents tracking and input voltages on the motors are shown in Figures 5, 6, 8, and 9. The simulation results show that the trajectory and force tracking errors asymptotically tend to zero, which validate the effectiveness of the control law in Theorem 5.1.

7. Conclusion

In this paper, effective robust control strategies have been presented systematically to control the holonomic constrained nonholonomic mobile manipulator in the presence of uncertainties and disturbances, and actuator dynamics is considered in the robust control. All control strategies have been designed to drive the system motion converge to the desired manifold and at the same time guarantee the boundedness of the constrained force. The proposed controls are nonregressor based and require no information on the system dynamics. Simulation studies have verified the effectiveness of the proposed controller.

Acknowledgment

The authors are thankful to the Ministry of Science and Technology of China as the paper is partially sponsored by the National High-Technology Research and Development Program of China (863 Program) (no. 2009AA012201).