5

Mobile Robot Control I

The Lyapunov-Based Method

This chapter deals with the general problem of determining the forces and torques that must be developed by the robotic actuators in order for the robot to go at a desired position/posture, track a desired trajectory, and, in general, to perform some task with a desired quality of performance. The controllers to be presented in the chapter assume that the goal of the control and the robot dynamic parameters are precisely known, and if this goal (posture or path) is changing, the change is compatible with the environment, and risk free. The particular objectives of the chapter are (i) to provide a minimal set of general control concepts and methods that are used in robot control, (ii) to study the basic general robot controllers that are applicable to all types of robots, and (iii) to present a number of feedback controllers, designed using the Lyapunov-based control theory, for the differential drive, car-like, and omnidirectional mobile robots. These controllers refer to the problems of position (posture) tracking, trajectory tracking, parking, and leader following.

Keywords

Lyapunov stability; DC motor model; state feedback control; computed torque control; kinematic tracking control; dynamic tracking control; polar coordinate-based robot model; parking control; leader–follower control; kinematic controller; dynamic controller; omnidirectional robot control

5.1 Introduction

Robot control deals with the problem of determining the forces and torques that must be developed by the robotic actuators in order for the robot to go at a desired position, track a desired trajectory, and, in general, to perform some task with desired performance requirements. The solution to control problems in robotics (fixed and mobile) is more complicated than usual due to the inertial forces, coupling reaction forces, and gravity effects. The performance requirements concern both the transient period and the steady-state period. In well-structured and fixed environments, such as the factory, the environment can be arranged to match the capabilities of the robot. In these cases, it can be assured that the robot knows certainly the configuration of the environment, and people are protected from the robot’s operation. In such controlled environments, it is sufficient to employ some type of model-based control, but in uncertain and varying (uncontrolled) environments, the control algorithms must be more sophisticated involving some kind of intelligence. The techniques to be presented in this chapter assume that the goal of the control and the robot kinematic and dynamic parameters are precisely known, and if this goal (posture or path) is changing, the change is compatible with the environment, and risk free.

Specifically, the objectives of the chapter are as follows:

• To provide a minimal set of general control concepts and methods that are used in the control of robots

• To study the basic general robot controllers that are applicable to all types of robots

• To present a number of feedback controllers, designed using the Lyapunov-based control theory, for the differential drive, car-like, and omnidirectional mobile robots.

These controllers refer to the problems of position (posture) tracking, trajectory tracking, parking, and leader following. In all cases, the control design involves two stages, viz., kinematic control (where only the kinematic models are used), and dynamic control (where the robot dynamics and actuators are also taken into account).

5.2 Background Concepts

In this section, the following fundamental control concepts and techniques are briefly discussed:

• State-space model

• Lyapunov stability

• State feedback control

• Second-order systems

Knowledge of these concepts is a basic prerequisite for the understanding of the material presented in the chapter. Full accounts are given in standard control textbooks [1].

5.2.1 State-Space Model

The state-space model of a control system is based on the concept of state vector image, which is the minimum dimensionality Euclidean vector, with components called state variables, the knowledge of which at an initial time image, together with the input vector image, for image, determines completely the behavior of the system for any time image. The dimension image of the state vector specifies the system’s dimensionality.

The above definition of the state means that the state of the system is determined by its initial value image at image and the input for image, and is independent of the state and the inputs for times previous to image.

It is noted that the state variables image of an n-dimensional system may not necessarily be measurable physical quantities, although in practice, an effort is made to use as more as possible measurable variables, because the state feedback control laws need all of them.

The expression of image as a function of image, and image, image, that is, image, is called the system’s trajectory.

The output image of the system is a similar function of image, image, and image, that is:

image

The trajectories satisfy the transition property:

image

for all image, where image.

In state-space model, the dynamic model of a nonlinear system (in continuous time) has the form:

image (5.1a)

image (5.1b)

where image and image are nonlinear vector functions of their arguments with proper dimensionality and the continuity and smoothness properties required in each case. The state vector image belongs to the state space image, the input (control) image belongs to the input space image, and the output image to the output space image, where image, image, image, and image is the n-dimensional Euclidean space.

If the vector functions image and image are linear, then the system is linear and is described by the model:

image (5.2a)

image (5.2b)

where image, image, image, image may be time-invariant or time-varying matrices of proper dimensionality (in many cases, image). A linear state-space model with the following matrices, image, image, and image (scalar), is called the controllable canonical model of the system that represents:

image (5.3)

For a scalar output image, a linear time-invariant system described by the nth-order differential equation:

image

or transfer function:

image (5.4)

where image is the complex frequency variable, can be modeled as in Eqs. (5.2a), (5.2b), and (5.3), if we define the state variables image as:

image (5.5)

Indeed, using Eq. (5.5) we get:

image

image

which gives the state-space model ((5.2a), (5.2b), and (5.3)) with:

image (5.6)

This model, also called phase variables canonical model, is very convenient for the pole-placement (or assignment) state feedback controller design.

The block diagram representation of the general model (5.2a) and ((5.2b)) has the form shown in Figure 5.1.

image

Figure 5.1 Block diagram of a general linear state-space model.

Other state-space canonical models of the system (5.2a) and ((5.2b)) are the observable canonical form and the Jordan canonical form fully described in control textbooks. To convert a given model (5.2a) and ((5.2b)) to some canonical form, use is made of a proper nonsingular linear (similarity) transformation image, where image is the new state vector.

Example 5.1

In this example, we derive the dynamic models (transfer function, state-space model) of the direct current (DC) electrical motor which is used in mobile robots to provide the torques that lead to the desired acceleration and velocity of them. DC motors are distinguished into motors controlled by the rotor (armature controlled), and motors controlled by the stator (field controlled). Both the motors will be considered.

Armature Controlled DC Motor

The rotor involves the armature and the commutator. A schematic of this motor is shown in Figure 5.2, where image and image are the resistance and inductance of the rotor, image is the load moment of inertia, and image is the linear friction coefficient.

image

Figure 5.2 Schematic of the armature controlled DC motor (image=rotation angle, image=motor angular speed, image=load moment of inertia).

The mechanical torque is given by:

image (5.7)

where image is the motor’s torque constant. The back electromotive force (emf) image which is subtracted from the input voltage image is proportional to image, that is:

image (5.8)

The characteristic curves of “image” are obtained by plotting the following function:

image

and have the linear form shown in Figure 5.3.

image

Figure 5.3 Characteristic curves of armature controlled system.

At the point where image, the motor maintains a constant angular speed (assuming of course that no external disturbances affect the motor). The differential equation of the motor can be found using the following relations:

image

where image, image is the moment of inertia of the load (plus the moment of inertia of the motor), and image is the linear friction coefficient, and eliminating image. The result is:

image

where imageimageimage imageimage. If image is negligible, then image, and the above differential equation reduces to:

image (5.9)

where image is as above, and image. In practice, the model in Eq. (5.9) represents an adequate approximation of the motor dynamics. The controllable state-space canonical model of the motor is found from Eq. (5.9) using Eqs. (5.3) and (5.6), that is:

image (5.10)

with image and image. A DC motor is shown in Figure 5.4A. Figure 5.4B shows DC motors of several sizes.

image

Figure 5.4 (A) A heavy-duty DC robot motor. (B) Several DC motors with embedded gear boxes. Source: http://www.goldmine-elec-products.com/prodinfo.asp?number=0; http://www.robojrr.tripod.com/motortech.htm.

Field Controlled DC motor

In this motor, the rotor’s current is kept constant (coming from a constant-current source), whereas the error is fed to the magnetic field of the stator via a phase-sensitive amplifier (Figure 5.5). The characteristic curves image show that when image is not excessively large, the mechanical torque image is independent of image, depending only on the field current image (Figure 5.5B).

image

Figure 5.5 (A) Field controlled DC motor. (B) Characteristic curves for different image (for very large image, we have a fall of image).

A schematic of this motor is shown in Figure 5.6.

image

Figure 5.6 Schematic of the field controlled DC motor.

Here, we have the following relations:

image (5.11)

where image is the field current/torque constant and image is the inductance of the stator, and all other symbols have the same meaning as in the armature controlled motor.

Combining the above equations we get:

image (5.12)

where image (DC gain), and image, image are the time constants of the magnetic field and the mechanical time constant of the motor, respectively, given by:

image

In practice, image is much smaller than image, and so Eq. (5.12) reduces to:

image (5.13)

which is of the same form as Eq. (5.9). Therefore, the motor has the state-space model (similar to Eq. (5.10)):

image

with image and image. If the motor dynamics (Eq. (5.9) or (5.13)) is expressed in terms of the angular velocity image, then (whenever the motor dynamics is not neglected) we use the form:

image

where image or image. This is analogous to a first-order RC circuit with time constant image.

5.2.2 Lyapunov Stability

Stability is a binary property of a system, that is, a system cannot be simultaneously stable or not stable. However, a stable system is characterized by a degree or index that shows how much near to instability the system is (relative stability). A system is defined to be bounded-input bounded-output (BIBO) stable if any bounded input leads always to a bounded output. A linear time-invariant system is BIBO stable if and only if all the poles of its transfer function or the eigenvalues of the matrix image of its state-space model (5.2a) and ((5.2b)) lie strictly on the left-hand complex semiplane image. The matrix image with the above property is called a Hurwitz matrix. The Routh and Hurwitz algebraic criteria specify the conditions that the coefficients of the system’s characteristic polynomial must satisfy in order for the system to be stable. A first-order system image (with a real pole image) is stable if image, and has the impulse response:

image

Since image tends to zero as image, asymptotically, the system is said to be asymptotically stable. Furthermore, since the convergence is exponential, that is, according to image, this system is called exponentially stable. For a second-order system where the matrix image has the eigenvalues image, image, the impulse response has the form:

image

Since image as image, the system is asymptotically stable, and because image, the system is exponentially stable. The above results hold for any combination of first-order and second-order systems.

The study of stability of a system and its stabilization via state or output feedback are two of the central problems in control theory. But the Routh and Hurwitz stability criteria can only be used for time-invariant linear single-input single-output (SISO) systems.

Lyapunov’s stability method can also be applied to time-varying systems and to nonlinear systems. Lyapunov has introduced a generalized notion of energy (called Lyapunov function) and studied dynamic systems without external input. Combining Lyapunov’s theory with the concept of BIBO stability we can derive stability conditions for input-to-state stability (ISS).

Lyapunov has introduced two stability methods. The first method requires the availability of the system’s time response (i.e., the solution of the differential equations). The second method, also called direct Lyapunov method, does not require the knowledge of the system’s time response.

Definition 5.1

The equilibrium state image of the free system image is stable in the Lyapunov sense (L-stable) if for every initial time image and every real number image, there exists some number image as small as desired, that depends on image and image, such that: if image, then image for all image, where image denotes the norm of the vector image, that is, image.■

Theorem 5.1

The transition matrix image of a linear system is bounded by image for all image if and only if the equilibrium state image of image is L-stable.■

The bound of image of a linear system does not depend on image. In general, if the system stability (of any kind) does not depend on image, we say that we have global (total) stability or stability in the large. If the stability depends on image, then it is called local stability. Clearly, total stability of a linear system implies also local stability.

Definition 5.2

The equilibrium state image is asymptotically stable if:

(i) It is L-stable.

(ii) For every image and image sufficiently near to image, the condition image, for image holds.■

Definition 5.3

If the parameter image in Definition 5.1 does not depend on image, then we have uniform L-stability.■

Definition 5.4

If the system image is uniformly L-stable, and for all image and for arbitrarily large image, the relation image implies image for image, then the system is called uniformly asymptotically stable.■

Theorem 5.2

The linear system image is uniformly asymptotically stable if and only if there exist two constant parameters image and image such that: image for all image and all image.■

Definition 5.5

The equilibrium state image of image is said to be unstable if for some real number image, some image and any real number image arbitrarily small, there always exists an initial state image such that image for image.■

Figure 5.7 illustrates geometrically the concepts of L-stability, L-asymptotic stability, and instability.

image

Figure 5.7 Illustration of L-stability (A), L-asymptotic stability (B), and instability (C). image and image symbolize n-dimensional balls (spheres) with radii image and image, respectively.

Direct Lyapunov method: Let image be the distance of the state image from the origin image (defined using any valid norm). If we find some distance image which tends to zero for image, then we conclude that the system is asymptotically stable. To show that a system is asymptotically stable using Lyapunov’s direct method, we do not need to find such a distance (norm), but a Lyapunov function which is actually a generalized energy function.

Definition 5.6

Time-invariant Lyapunov function is called any scalar function image of image which for all image and image in the vicinity of the origin satisfies the following four conditions:1

(i) image is continuous and has continuous derivatives

(ii) image

(iii) image for all image

(iv) image

Theorem 5.3

If a Lyapunov function image can be found for the state of a nonlinear or linear system image, where image (image is a general function), then the state image is asymptotically stable.■

Remarks

(i) If Definition 5.6 holds for all image, then we have “uniformly asymptotic stability.”

(ii) If the system is linear, or we replace in Definition 5.6 the condition “image in the vicinity of the origin,” by the condition “image everywhere,” then we have “total asymptotic stability.”

(iii) If the condition (iv) of Definition 5.6 becomes image, then we have (simple) L-stability.

Clearly, to establish L-stability of a system, we must find a Lyapunov function. Unfortunately, there does not exist a general methodology for this.

Analogous results hold for the case of time-varying Lyapunov functions image, namely:

Definition 5.7

Time-varying Lyapunov function image for the system’s state is any scalar function of image and image, which, for all image and image near the origin image, has the following properties:

(i) image and its partial derivatives exist and are continuous

(ii) image

(iii) image for image and image, where image and image is a scalar continuous non decreasing function of image

(iv) image

Remark

For all image, the function image should have values greater than or equal to the values of some continuous nonreducing time-invariant function.

Theorem 5.4

If a time-varying Lyapunov function image can be found for the state image of the system image, then the state image is asymptotically stable.■

Definition 5.8

(i) If the conditions of Definition 5.7 hold for all image and image, where image is a continuous nondecreasing scalar function of image with image, then we have uniformly asymptotic stability.

(ii) If the system is linear or the conditions of Definition 5.7 hold everywhere (not only in a region of the origin image), and if we have image, when image, then we say that the system is uniformly totally asymptotically stable.■

In the case of a linear time-varying system image the Lyapunov (time varying) function image is given by the quadratic (energy) function:

image (5.14a)

where image satisfies the following matrix differential equation:

image (5.14b)

If the system is time-invariant image, then image and the above differential equation for image reduces to the following algebraic equation for image:

image (5.15)

In this case, we can select a positive definite matrix image and solve the image equations (image is symmetric) for the elements of image. Then, if image (i.e., if image is positive definite) the system is asymptotically stable.

Remark

Using Eq. (5.15), we can show that the Lyapunov stability criterion is equivalent to the Hurwitz stability criterion.

5.2.3 State Feedback Control

State feedback control is more powerful than classical control because the design of a total controller for a multiple-input multiple-output (MIMO) system is performed in a unified way for all control loops simultaneously, and not serially one loop after the other which does not guarantee the overall system stability and robustness. In this section we will briefly review the eigenvalue placement controller for SISO systems.

Let a SISO system:

image

where image is a image constant matrix, image is an image constant matrix (column vector), image is an image matrix (row vector), image is a scalar input, and image is a scalar constant. In this case, a state feedback controller has the form:

image (5.16)

where image is a new control input and image is an n-dimensional constant row vector: image. Introducing this control law into the system, we get the state equations of the closed-loop (feedback) system:

image (5.17)

The eigenvalue placement design problem is to select the controller gain matrix image such that the eigenvalues of the closed-loop matrix image are placed at desired positions image. It can be shown that this can be done (i.e., the system eigenvalues are controllable by state feedback) if and only if the system image is totally controllable. The concept of controllability has been developed to study the ability of a controller to alter the performance of the system in an arbitrary desired way. As it is known, the positions of the eigenvalues specify the performance characteristics of a system.

Definition 5.9

A state image of a system is called totally controllable if it can be driven to a final state image as quickly as desired independently of the initial time image. A system is said to be totally controllable if all of its states are totally controllable.■

Intuitively, we can see that if some state variables do not depend on the control input image, no way exists that can drive it to some other desired state. Thus, this state is called a noncontrollable state. If a system has at least one noncontrollable state, it is said to be nontotally controllable or, simply, noncontrollable. The above controllability concept refers to the states of a system and so it is characterized as state controllability. If the controllability is referred to the outputs of a system then we have the so-called output controllability. In general, state controllability and output controllability are not the same.

Theorem 5.5

The necessary and sufficient condition for a linear system image to be totally state controllable is that the controllability matrix:

image (5.18a)

has

image (5.18b)

where image is the dimensionality of the state vector image.■

The most straightforward technique for selecting the feedback matrix image is through the use of the controllable canonical form. This technique involves the following steps:

Step 1: We write down the characteristic polynomial image of the matrix image:

image

Step 2: Then, we find a similarity transformation image that converts the given system to its controllable canonical form image.

Step 3: From the desired eigenvalues of the closed-loop system, we determine the desired characteristic polynomial:

image


The feedback gain matrix image of the controllable canonical model is given by:

image

Step 4: Equating the last rows of image and image we find:

image


and so solving for image we find:

image (5.19)

5.2.4 Second-Order Systems

The state vector image of a second-order system contains the position and velocity of the variable (physical quantity) of interest, namely:

image

Suppose that it is desired to design the state feedback controller such that the system’s state follows a desired trajectory image. In this case, the feedback must use the measured error:2

image (5.20)

and the controller should reduce the sensitivity of the system to the inaccuracy and the uncertainty in the parameter values used in the dynamic model.

A fundamental characteristic of control systems is the bandwidth image which determines the operation speed of a system and capability of fast trajectory tracking. The greater the bandwidth is the better, but image should not be very high because it may excite possible high-frequency components that have not been included in the system model.

As an example, consider the following simple second-order system:

image (5.21)

which by Eq. (5.20) gives the error equation:

image (5.22)

Defining the state vector:

image

we get the following controllable canonical form:

image (5.23a)

where:

image (5.23b)

To get a desired bandwidth image, we select the desired closed-loop characteristic polynomial as:

image (5.24)

where image is the undamped natural frequency (taken equal to the desired bandwidth image) and image is the damping coefficient (usually selected image).

Then, the feedback controller is selected as in Eqs. (5.16) and (5.19) with image (unit matrix), namely:

image

or

image (5.25)

The closed loop error system is obtained using Eqs. (5.22) and (5.25), that is:

image (5.26)

which has the desired damping and bandwidth specifications.

The control law (5.25) contains the proportional term image and the derivative term image, that is, it is a PD (proportional plus derivative controller) which is one of the most popular and efficient controllers. For second-order systems, the PD controller gives exact results.

Example 5.2

Consider the case where the system (5.21) is corrupted by a disturbance image as:

image

(a) It is desired to determine the steady-state position error image of the closed-loop PD controlled system when image is a step disturbance of amplitude image:

image

(b) Show that this steady-state error vanishes if, instead of the PD controller, a PID (proportional plus integral plus derivative) controller is used.

Solution

(a) With the PD controller, the closed-loop disturbed system (5.26) becomes:

image

With image and image, the above step disturbance, the steady-state position error, obtained by setting image and image, satisfies the relation:

image

Thus:

image

We see that image has a nonzero finite value which is proportional to image and inversely proportional to the square of the bandwidth image.

(b) If we use a PID controller:

image

the closed-loop error system becomes:

image

Then, in the limit as image, for image, image and image, image we get:

image

which implies image. Otherwise, image cannot have a finite value.

Remark

The above results can also be obtained using the well-known Laplace transform final-value property image and computing image via the error system’s transfer function.

Example 5.3

It is desired to check the stability of the following systems using the Lyapunov method:

(a) image

(b) image
image

Solution

System (a): This system has the equilibrium state image. We examine if the function image is a Lyapunov function. This is done by checking if all conditions for image to be a Lyapunov function are satisfied. We have:

• image and image are continuous

• image

• image for image

• image for all image

We see that all properties of the candidate function required to be a Lyapunov function hold, and so the system image is uniformly asymptotically stable.

System (b): We try the following candidate Lyapunov function:

image

This function has the following properties:

• image and image are continuous

• image

• image for image

• image

that is, it possesses all properties of a Lyapunov function. Therefore, the only equilibrium state image (i.e., the origin) is totally asymptotically stable.

5.3 General Robot Controllers

The following general controllers, which are standard in robotics [24], will be examined:

• Proportional plus derivative control

• Lyapunov function based control

• Computed torque control

• Resolved motion rate control

• Resolved motion acceleration control

5.3.1 Proportional Plus Derivative Position Control

Here, it will be shown that PD control leads to satisfactory results in the control of the position of a general robot described in Eqs. (3.11a) and (3.11b):

image

image

where for any image, image is a known positive definite matrix. Assuming that the friction is negligible and omitting the gravity term image, which anyway is zero in mobile robots moving on an horizontal terrain, we get:

image (5.27)

where image is defined as in Eq. (3.13), and the matrix image is antisymmetric.

Let image be the error between image and image. Then, the PD controller has the form

image (5.28)

where image and image are positive definite symmetric matrices. The resulting feedback control scheme has the form of Figure 5.8.

image

Figure 5.8 PD robot control.

Let us try the following candidate Lyapunov function:

image (5.29)

where the term image is the robot’s kinetic energy, and the term image represents the proportional control term. Thus, the function image can be considered as representing the total energy of the closed-loop system. Since image and image are symmetric positive definite matrices, we have image and image for image. Therefore, we have to check the validity of property (iv) of Definition 5.6.

Here, Eq. (5.29) gives:

image (5.30)

Now, introducing the control law (5.28) into Eq. (5.30) we get:

image (5.31)

Therefore, since the matrix image is antisymmetric, Eq. (5.31) finally gives:

image (5.32)

We observe that while the Lyapunov function image in Eq. (5.29) depends on image, its derivative image depends on image, which is analogous of the known property of classical SISO PD control. The condition (5.32) ensures that the feedback error control system (Figure 5.8) is L-stable. It is also useful to remark that the PD control (5.28) is particularly robust with respect to mass variations because it does not require knowledge of the parameters that depend on mass.

A special case of the controller (5.28) is:

image

which is applied to each joint separately. If the motion is subject to friction (assumed linear) the robot model (5.27) must simply be replaced by:

image

where image is the diagonal matrix of friction coefficients. The present PD controller can be enhanced with an integral term as given in Example 5.2.

5.3.2 Lyapunov Stability-Based Control Design

The control design method applied to the above problem is known as Lyapunov-based controller design and constitutes a widely used method for both linear and nonlinear systems. The steps of this method are the following:

Step 1: Select a trial (candidate) Lyapunov function, which is typically some kind of energy-like function for the system, and possesses the first three properties of Lyapunov functions (Definition 5.6, Section 5.2.2).

Step 2: Derive the equation for the derivative image along the system trajectory:

image


and select a feedback control law:

image


which ensures that:

image


Typically, image is a nonlinear function of image that contains some parameters and gains which can be selected to make image, and thus ensures that the closed-loop system is asymptotically stable.

Remark

For a given system, one may find several Lyapunov functions and corresponding stabilization controllers. If the system is linear time-invariant image, then it is not necessary to work using the Lyapunov method. In this case, the controller is a static linear state feedback controller image (Eq. (5.16)), which can be selected to make the closed-loop matrix a Hurwitz matrix (with all its eigenvalues on the strict left-hand s-plane). This assures that the closed-loop system is asymptotically and exponentially stable. The Lyapunov-based controller design method will be applied as a rule in most cases of the discussions that follow.

5.3.3 Computed Torque Control

The computed torque control technique reduces the effects of the uncertainty in all the terms of the Lagrange model. The controller image is selected to have the same form as the dynamic model (Eq. (3.11a)), that is:

image (5.33)

Thus, since the inertia matrix is positive definite (and so invertible), introducing the control law (5.33) in the system (3.11a), we get:

image (5.34)

This implies that image can be a decoupled controller (PD, PID) that can control each joint (motor axis) independently. The basic problem of the computed torque method is that we do not have available the exact values of image, image, and image, but only approximate values image, image, and image. Then, instead of Eq. (5.33) we get:

image (5.35)

and so Eq. (5.34) is replaced by:

image (5.36)

A problem with the model in Eq. (5.35) is the investigation of its robustness to modeling uncertainties which include uncertainties in the parameter values and nonmodeled high-frequency components (e.g., structural resonance, sampling rate, or omitted time delays). The computed torque control method belongs to the general class of linearization techniques via nonlinear state feedback (see Section 6.3). Solving the model (Eq. (3.11a)) for image we get:

image (5.37)

Introducing the controller (5.35) into (5.37) we obtain the block diagram of the overall closed-loop system shown in Figure 5.9.

image

Figure 5.9 Closed-loop computed torque control system.

5.3.4 Robot Control in Cartesian Space

The controllers presented thus far work in the joints’ (motors’) space and are based on the error image between the actual and desired generalized joint variables image (called internal variables). The motion of the robot in the Cartesian (or task, or working) space is obtained indirectly from the motion of the joints. However, in many cases, it is required to design the controller so as to work directly with the Cartesian variables, called external variables.

In Cartesian space, we have three types of controllers which are known as resolved motion controllers:

• Resolved motion rate control

• Resolved motion acceleration control

• Resolved force control

Here, we will study the resolved motion rate control which is mostly used in mobile robots. The resolved acceleration controller is actually an extension of the resolved motion rate control that includes the acceleration, a fact that will also be briefly considered.

5.3.4.1 Resolved Motion Rate Control

The resolved motion rate control is the control where the joints are moved simultaneously in different velocities, such that a desired motion in Cartesian (or task) space is obtained.

In general, the relation of the linear and angular velocity (motion rate) vector:

image

in Cartesian space and the velocities image of the robotic joints is given by the Jacobian relation (Eq. (2.6)):

image (5.38)

where the Jacobian matrix image is given by Eq. (2.5), which has the inverse (see Eqs. (2.7b) and (2.8)):

image (5.39a)

or the generalized inverse:

image (5.39b)

Differentiating Eq. (5.38) we get:

image (5.40)

Thus, introducing into Eq. (5.40) the expression of image given by Eq. (5.39a) yields:

image (5.41)

This relation gives the accelerations of the joints for given linear/angular velocity and acceleration of the robot end-effector in Cartesian space. If image is not square, we use the generalized inverse image in place of image.

The block diagram of the resolved velocity control is shown in Figure 5.10.

image

Figure 5.10 Block diagram of the robot resolved rate control based on Eq. (5.39a).

In the simplest case, the joints control can be a proportional control law with gain image. In many cases, the task space control is required to be in a coordinate frame attached to the robot (and not in the world coordinate frame). The velocity image is then given by:

image (5.42)

where image is the desired velocity of the robot and image is a matrix that relates image to image. Then, from Eqs. (5.39a) and (5.39b) we get:

image (5.43)

The relation (5.43) is typically used in vision-based robot control (visual servoing).

5.3.4.2 Resolved Motion Acceleration Control

The resolved motion acceleration control method is based on the equation:

image (5.44)

which is found by differentiating Eq. (5.38). The desired position, velocity, and acceleration of the robot in Cartesian space are assumed to be known from the trajectory planner. Thus, to reduce the position error, we must apply appropriate forces/torques at the robot joints, such that the acceleration in Cartesian space satisfies the relation:

image (5.45)

where image, and image are the desired translation position, velocity, and acceleration, respectively.

Here, the position error is:

image

Therefore, in terms of image, Eq. (5.45) is written as:

image (5.46)

and the gains image should be selected such that image tends asymptotically to zero. A similar error equation can be derived for the angular acceleration image, using the control law:

image (5.47)

where image is the orientation angle in the Cartesian space. Combining Eqs. (5.45) and (5.47), and inserting into Eq. (5.44), we get:

image (5.48)

where:

image

Equation (5.48) constitutes the basis for the resolved motion acceleration control of robots. The position image and velocity image are measured by potentiometers or optical encoders.

5.4 Control of Differential Drive Mobile Robot

The control procedure will involve two stages:

1. Kinematic stabilizing control

2. Dynamic stabilizing control

The resulting linear and angular velocities in the kinematic stage will be used as reference inputs for the dynamic stage. Thus, this procedure belongs to the general class of backstepping control [513].

5.4.1 Nonlinear Kinematic Tracking Control

The robot motion is governed by the dynamic model (Eqs. (3.23a) and (3.23b), the kinematic model (Eq. (2.26)), and the nonholonomic constraint (Eq. (2.27)), namely:

image (5.49a)

image (5.49b)

image (5.49c)

where, for notational simplicity, the index image was dropped from image, image and image, image, image are the control inputs, with:

image (5.49d)

being the state vector.

The problem is to track a desired state trajectory:

image (5.50)

with error that goes asymptotically to zero.

To this end, the Lyapunov stabilizing method will be used. For realizability, the desired trajectory must satisfy both the kinematic equations and the nonholonomic constraint, that is:3

image (5.51)

The errors image, image, and image, expressed in the wheeled mobile robots’ (WMR’s) local (moving) coordinate frame image, are given by (see Eq. (2.17)):

image (5.52a)

and

image (5.52b)

Differentiating Eqs. (5.52a) and (5.52b) and taking into account Eqs. (5.49c) and (5.51), we get the following kinematic model for the error image:

image (5.53)

where the linear and angular velocities image and image are the kinematic control variables. Clearly, Eq. (5.53) satisfies the kinematic and nonholonomic equations of the WMR.

Therefore, the kinematic feedback controller will be based on Eq. (5.53). The Lyapunov stabilizing method of Section 5.3.2 will be applied. Since here the controller should be nonlinear, we cannot select its structure beforehand. Its structure will be determined by the choice of the candidate Lyapunov function. Here, the following candidate function is selected [14]:

image (5.54)

This function satisfies the first three properties of Lyapunov functions, namely:

(i) image is continuous and has continuous derivatives

(ii) image

(iii) image for all image

We therefore have to check under what conditions the fourth property can be satisfied.

Differentiating Eq. (5.54) with respect to time we get:

image (5.55)

To make image, the control inputs image and image are selected such that:

image (5.56)

which leads to:

image (5.57a)

image (5.57b)

Clearly, for image and image, we have image with the equality obtained only when image and image. Thus, the controller (5.57a) and ((5.57b) guarantees total asymptotic tracking to the desired trajectory.

Remark 1

We can also add a gain image in the second term of Eq. (5.56b), that is, choose image as

image (5.58a)

In this case, total asymptotic tracking to the desired trajectory image can be proved using the following Lyapunov function:

image (5.58b)

Remark 2

A more general kinematic controller can be obtained by using the Lyapunov function:

image (5.58c)

with image, image, and image. In the following, we will derive this controller. To this end, we define new control variables image, image and write the model (5.53) as:

image (5.58d)

Differentiating Eq. (5.58c) and using the model (5.58d) we get:

image

To assure that image, we use two functions image and image for all image, and select image and image such that:

image (5.58e)

Then, it follows that:

image

The above controller assures that image and imageimage. Since image is uniformly continuous, Barbalat’s Lemma (Sec. 6.2.3) implies that image. This by Eq. (5.58e) implies that image and imageimage. Now, obviously, image (i.e., image). One way to overcome the fact that image does not converge only to 0 but also to image is to take care that the WMR, before trying to track immediately the desired trajectory (i.e., the virtual WMR), is rotating about its own axis with an increasing angular velocity image until it sees the virtual robot. The reader can verify that this can be done by the controller [11]:

image

where image is given by the dynamic model:

image

with image being a step input function defined by:

image

5.4.2 Dynamic Tracking Control

Having selected image and image as in Eqs. (5.57a) and (5.57b) (or Eq. (5.58a)), we select the control inputs (torques) image and image in Eq. (5.49a) as:

image (5.59a)

image (5.59b)

where:

image (5.59c)

Introducing Eqs. (5.59a)(5.59c) into Eqs. (5.49a) and (5.49b) we get the velocities’ error equations:

image

which for image and image are stable and image, image converge to zero asymptotically. Therefore, in selecting the feedback control inputs (torques) as in Eqs. (5.59a) and (5.59b) with image and image given by Eqs. (5.57a) and (5.57b), the tracking of the desired trajectory image is achieved asymptotically, as required. The block diagram of the feedback tracking controller is depicted in Figure 5.11.

image

Figure 5.11 The complete trajectory tracking feedback control system of the WMR.

5.5 Computed Torque Control of Differential Drive Mobile Robot

The control design procedure involves again two stages: kinematic control followed by dynamic control. We will work on the WMR shown in Figure 5.12, where the motor dynamics includes a gear box (of ratio N) [13].

image

Figure 5.12 Differential drive WMR where the point image traced by the controller is different than image and image.

Here, Q is the midpoint of wheel baseline, G the center of gravity, and C the point traced by the controller (different than the point Q).

The meaning of the remaining symbols are self-evident (the same as in Figure 2.7).

5.5.1 Kinematic Tracking Control

The kinematic equations of the robot are:

image (5.60a)

image (5.60b)

image (5.60c)

where image is the lateral velocity in the local coordinate frame image.

Equations (5.60a) and (5.60b) are written in the matrix form:

image

Now, using new control variables image and image defined by:

image (5.61)

we get:

image (5.62)

The dynamic system (5.62) is linear and decoupled, and so the state-feedback law:

image (5.63)

yields the error dynamics:

image (5.64)

with image and image.

Therefore, from Eq. (5.64), it follows that for any positive gains:

image

the tracking error tends exponentially to zero.

Combining Eqs. (5.61) and (5.63) we get the overall nonlinear kinematic control law:

image (5.65)

5.5.2 Dynamic Tracking Control

The feedback kinematic controller (5.65) incorporates the WMR kinematic equations and so one can now use the reduced (unconstrained) dynamic model ((3.19a) and (3.19b)) of the robot for the selection of the control inputs (motor torques or motor voltages), as described in Section 5.4.2, where actually the computed torque method was applied.

For the robot of Figure 5.12, with the motor dynamics included, the reduced model has the following form:

image (5.66)

where:

image (5.67)

with:

image (5.68)

Here:

image=combined wheel, motor rotor, and gearbox inertia

image=combined wheel, motor, and gearbox friction coefficient

image=right and left wheel motor voltage

image=electrical resistance

image=motor voltage/torque constants

Now, applying the computed torque (linearization) technique to Eq. (5.66), we choose the voltage control vector image as:

image (5.69)

where image is the new control vector. Introducing Eq. (5.69) into Eq. (5.66) we get:

image

with:

image

Therefore, selecting the linear state feedback control law:

image (5.70)

yields the error system:

image

which for image is asymptotically stable with equilibrium state image.

Combining Eq. (5.69) with Eq. (5.70), we get the full dynamic controller:

image (5.71)

The overall tracking controller of the robot is given by Eqs. (5.65) and (5.71).

Example 5.4

It is desired to formulate and solve the kinematic problem of a unicycle-like WMR to go asymptotically from an initial state (position and orientation) to a goal state (position and orientation) using polar coordinates.

Solution

The unicycle-like robots are described by the kinematic Eq. (2.26):

image (5.72)

where image is the velocity and image is the angle of the WMR image axis with the goal (world) coordinate frame x-axis. All WMRs with the above kinematic equations, for example, the differential drive WMR, are said to belong to the unicycle-like class of mobile robots. The geometry of the WMR that will be used for formulating the polar coordinates is shown in Figure 5.13 [15].

image

Figure 5.13 Polar coordinates of the unicycle. Here, the goal coordinate frame Gxy is considered to be the world coordinate frame.

The kinematic control variables of the robot are image and image. The polar coordinates (position and orientation) of the WMR are its distance image from the goal, and its orientation image with respect to the goal’s coordinate frame Gxy. The steering angle is image. Then, in polar coordinates, the kinematic model in Eq. (5.72) is replaced by:

image (5.73a)

image (5.73b)

image (5.73c)

These relations hold for image, a condition which will be always satisfied by an asymptotic reduction of image to zero (since for any finite time there always be image).

Our goal tracking control problem is to find a state-feedback law:

image (5.74)

which guarantees that image, image, and image, asymptotically.

To this end, we will apply the Lyapunov-based control method. Let us choose the following candidate Lyapunov function [15]:

image

image (5.75)

Clearly, the function image possesses the first three properties of Lyapunov functions. We will determine the controller (5.74) which will assure that the fourth property image is also possessed by image, along the system trajectory.

The time derivative of image, along the trajectory determined by Eqs. (5.73a)(5.73c), is found to be:

image (5.76a)

where:

image (5.76b)

image (5.76c)

Choosing image as:

image (5.77)

yields:

image (5.78)

Now, introducing Eq. (5.77) into Eq. (5.76c) we get:

image

Thus, selecting image as:

image (5.79)

yields:

image (5.80)

The inequalities (5.78) and (5.80) give:

image (5.81)

This implies, by the Lyapunov stability theorem, that image and image go asymptotically to zero for any image. Thus, we have to see what happens to image. To this end, we get the closed-loop system kinematics by introducing the control laws (5.77) and ((5.79)) into Eqs. (5.73a)(5.73c), namely:

image (5.82a)

image (5.82b)

image (5.82c)

These equations show that the asymptotic convergence of image and image to zero implies the asymptotic convergence of image to its only equilibrium state image. In fact, from Eq. (5.82c) it follows that image implies image, that is, image tends to some finite value image. Then, we see from Eq. (5.82b) that the uniformly continuous function image tends necessarily to image.4 On the other hand, by Barbalat’s lemma, image tends to zero, which in turn implies that image (see Section 6.2.3). Therefore, the smooth kinematic control laws (5.77) and ((5.79)) assure the asymptotic tracking of the goal position and orientation by the WMR, as desired. The fact that the control laws (5.77) and ((5.79)) are continuously differentiable does not contradict Brockett’s Theorem 6.6 and its corollary (c) because this theorem is valid for the Cartesian state-space representation (Eq. (5.72)) of the unicycle-like WMR [15].


4Recall that image as image.

5.6 Car-Like Mobile Robot Control

For the car-like WMR, we will study the following two representative problems:

1. Parking (or posture) control

2. Leader–follower (formation) control

5.6.1 Parking Control

Consider a car-like WMR (Figure 2.9) which controls the steering angle image and the rear-wheels’ velocity image. The orientation of the car body (i.e., of image) is image. The kinematic equations of the robot are given by Eq. (2.52):

image (5.83)

The problem is to control the WMR (using image and image) so as to move it to a desired parking position and orientation, which here is assumed to be image, image, and image (as it was actually done in Example 5.4 for the unicycle WMR). Here, this problem will be solved by a two-step maneuver to overcome the turning radius limitation of the car-like mobile robot, namely [16]:

Step 1: Controller stabilizing image and image

Step 2: Controller stabilizing image and image

Use of the Lyapunov-based control method will again be made as usual.

Step 1:imagecontrol

We select the following candidate Lyapunov function:

image

which satisfies the first three conditions of Lyapunov functions. We will check if the third condition can be satisfied, along the trajectory of the system in Eq. (5.83). We have:

image (5.84)

Choosing:

image (5.85)

and introducing into Eq. (5.84) yields:

image (5.86)

which, by Lyapunov theorem, implies that image tends asymptotically to zero.

The closed loop kinematics equation is:

image (5.87)

Let image. For image to stay zero, image should be zero. Then, Eq. (5.87) implies that image should also tend to zero, for image.

The change of the sign of image is needed when the WMR cannot move with its current velocity due to the presence of obstacles or when the measured states exceed some predetermined bounds. Of course, the initial selection of image affects the efficiency of the path. Actually, the controller (5.85) is a nonlinear bang-bang controller. Therefore, a switching rule is needed to determine when the change from image to image must occur.

Defining image, the switching rule is:

If image

Then image; otherwise image

This rule means that if the front part of the WMR is closer to the origin, then it will go forward or backward.

Step 2:imagecontrol

We use the same form of the candidate Lyapunov function:

image

with image and image.

The time derivative image along the trajectory of Eq. (5.83) is:

image

Choosing:

image (5.88)

gives:

image (5.89)

which is negative semidefinite for image.

For image, in which case, by Eq. (5.88), image, we have

image

Therefore, the mobile robot is uniformly L-stable at image. However, image cannot converge without increasing image, a fact which is due to the low bound of the WMR turning radius (Figure 5.14A) [16,17]. Actually, image can be made arbitrarily small at the first step. Therefore, when we achieve a very small image (i.e., image), we use image, and image.

image

Figure 5.14 (A) For image to converge, image must be increased. (B) Case image. (C) Case image. (D) Convergence to the origin image using Eq. (5.90).

This situation is overcome if we use the transformation [16]:

image (5.90)

In this way, the distance from the origin to the WMR is equal to the error of image, and the difference angle of the WMR with the x-axis becomes the error of image. As shown in Figure 5.14D, the controller (5.88) with the above switching type transformation assures that the WMR can go to image starting from any initial posture.

5.6.2 Leader–Follower Control

Consider two car-like robots that follow a path with the first car acting as the leader and the second being a follower (Figure 5.15). For more WMRs, one following the other in front of it, this problem is known as formation control [7].

image

Figure 5.15 (A) Two car-like WMRs (leader–follower structure). (B) Four WMRs in a typical formation (diamond structure). Source: www.robot.uji.es/lab/plone/research/pnebot/index_html2.

The leader–follower control problem under consideration is to find a velocity control input for the follower that assures convergence of the relative distance image and relative bearing angle image of the WMRs to their desired values, under the assumption that the leader motion is known and is the result of an independent control law [7]. To solve the problem, we will apply the Lyapunov-based control design method using the kinematic and dynamic equations of the bicycle equivalent presented in Section 3.4 (see Eqs. (3.56), (3.57), (3.60a)(3.60d)).

In Figure 5.15, image, image, and image are the linear velocity, orientation angle, and steering angle of the leader, and image, image, image are the respective variables of the follower. The coordinates of points image and image are denoted by image and image.

We first derive the dynamic equations for the errors:

image (5.91)

where image, image, and image represent the desired trajectory of the follower in world coordinates, which are transformed to image, image, and image in the local coordinate frame of the follower. From Figure 5.15 we get:

image (5.92a)

image (5.92b)

image (5.92c)

image (5.92d)

Differentiating Eqs. (5.92b) and (5.92c) gives:

image (5.93a)

image (5.93b)

with (see Eqs. (3.56) and (3.57)):

image (5.93c)

where image.

Using Eqs. (3.56) and (5.93c) in Eqs. (5.93a) and (5.93b), we obtain:

image (5.94a)

image (5.94b)

while from Figure 5.13 we have:

image (5.95)

Then, differentiating Eqs. (5.92a) and (5.92d), introducing Eqs. (5.94a) and (5.94b), and using the auxiliary variable:

image

we get, after some algebraic manipulation:

image (5.96a)

image (5.96b)

Using Figure 5.15, the actual and desired coordinates of the follower’s point A can be expressed in terms of the coordinates of the leader’s point B. Therefore, we use the variables image and image and get the error equations (in the world coordinate frame):

image (5.97a)

image (5.97b)

image (5.97c)

Finally, differentiating Eqs. (5.97a)(5.97c), we obtain the dynamic equations for image, image, and image:

image (5.98a)

image (5.98b)

image (5.98c)

We are now ready to apply the usual two-stage (kinematic, dynamic) backstep controller design.

5.6.2.1 Kinematic Controller

We select the candidate Lyapunov function [7]:

image (5.99)

which is similar to Eq. (5.54), which possesses the first three properties of Lyapunov functions. The feedback control inputs image and image will be selected such that to make image. Differentiating Eq. (5.99) gives:

image (5.100)

Introducing Eqs. (5.98a)(5.98c) into Eq. (5.100), we find that the selection of image and image as:

image (5.101a)

image (5.101b)

with image makes image. Indeed, introducing Eqs. (5.98a)(5.98c), (5.101a), and (5.101b) into Eq. (5.100) yields:

image

Since image, choosing image, image, and image makes image.

5.6.2.2 Dynamic Controller

Use will be made of the dynamic model (Eqs. (3.60a)–(3.60d)). The desired velocity and steering angle image, image of the follower are given by the results of the kinematic controller. We define the error:

image (5.102)

From Eqs. (3.60a)–(3.60c), applied to the follower WMR, we get:

image (5.103)

where image is the driving torque of the steering wheel of radius image, and the relation image was used. Combining Eqs. (5.103) and (3.60d) gives:

image (5.104)

where:

image

image

image

Subtracting both sides of Eq. (5.104) from image we get:

image

Now, adding and subtracting image to the right-hand side of this equation yields:

image (5.105a)

where:

image (5.105b)

and image:

The torque image in Eq. (5.105a) can be found using the computed torque technique as:

image (5.106)

which is introduced into Eq. (5.105a) to give the closed-loop error equation:

image (5.107)

It only remains to select image such that image tends to zero asymptotically. We select the candidate Lyapunov function:

image (5.108)

where image is given by Eq. (5.99).

Differentiating image and introducing the result into Eq. (5.107) gives:

image

Since image by the kinematic controller design, we can assure that image by selecting the gain matrix image such that the matrix image is positive definite, that is:

image

with image and image. Then, we find:

image

This implies that the error image in Eq. (5.105a) tends asymptotically to zero, that is, image and image, as required. The function image in the control law Eq. (5.106) can be approximated by a neural network using a weight updating rule as described in Section 8.5 [7].

5.7 Omnidirectional Mobile Robot Control

We will consider the three-wheel omnidirectional dynamic robot model (Eqs. (3.77a) and (3.77b)) derived in Section 3.5:

image (5.109)

and will apply the resolved motion acceleration technique, combined with PI or PD control [18]. Solving Eq. (5.109) for image, we get the inverse dynamic resolved acceleration equations:

image (5.110a)

image (5.110b)

image (5.110c)

where:

image (5.111a)

image (5.111b)

image (5.111c)

with:

image

being the errors between the desired and actual trajectories, and image, image, image being proportional gains, image, image integral gains, and image the derivative (velocity) gain of image. Note that the factors image and image in Eqs. (5.110a)(5.110c) are always nonzero (see Eqs. (3.76a)–(3.76c)) and so the above resolved acceleration controllers image, image, and image exist for all image.5

The block diagram of the overall feedback control WMR system is shown in Figure 5.16.

image

Figure 5.16 Resolved acceleration control system for the three-wheel omnidirectional robot (image represents the robot inverse dynamics Eqs. (5.110a)(5.110c)).

The above controller was applied to a real robot with physical parameters:

image

The initial state used is:

image

A basic experiment was to check the WMRs holonomic property, that is, the ability of the robot to independently achieve translational and rotational motion around the center of gravity in the x-y plane.

To check this property, it was assumed that the robot must travel with a single azimuth image (see Figure 3.6) for 20 s, but with zero rotational angle for 0–10 s, and a uniformly varying rotational angle from image to image for 10–20 s. Although the moving velocity was to be image in the steady state, a sinusoidal reference velocity was set for each 2 s at the starting and ending period. Using these image and image values, the corresponding image and image were derived from:

image

where the positive rotational direction of motion is the counterclockwise direction. The desired accelerations image and image were derived by differentiation of image and image. A selection of the gains used is given in Table 5.1.

Table 5.1

Gains of the Resolved Acceleration PI/PD Controller

Image

The responses obtained for image, image, and image, and the (x,y) trajectory are shown in Figure 5.17.

image

Figure 5.17 (A,C) Actual velocity responses compared to the desired velocity responses image and image. (B) Trajectory of image, (D) (x,y) trajectory. Source: Reprinted from Ref. [18], with permission from Springer Science+Business Media BV.

We observe that despite the occurrence of some oscillations in the velocities image and image, the image trajectory matches perfectly the desired trajectory image. Another experiment was carried out to follow a circular path with a radius less than half the distance between the wheels (say 0.1 m with distance between the wheels 0.356 m). Using the same gains as before, the results were very good, with perfect tracking of the circular path [18].

Example 5.5

Apply the computed torque technique to derive a PD path tracking controller for a three-wheel omnidirectional robot.

Solution

We will work with the dynamic model derived for the omnidirectional robot of Figure 3.7 in Example 3.2, namely:

image (5.112)

The computed torque control law has the form (Eq. (5.33)):

image (5.113)

and the PD control has the form:

image (5.114)

where:

image (5.115)

is the tracking error of the resulting path from the desired path image. Combining Eqs. (5.113) and (5.114) and introducing into Eq. (5.112), we get the closed-loop system:

image

or the tracking error dynamic equation:

image (5.116)

since the matrix image is nonsingular. We see that the origin image is an equilibrium point of this system. Therefore, selecting proper values of image and image, we can assure that image tends to zero, that is, image, asymptotically. To this end, we use the following candidate Lyapunov function:

image (5.117)

which, for sufficiently small constant image, satisfies the first two properties of Lyapunov functions.

We will now examine under what conditions image is negative. The time derivative of image along the trajectory of image is found to be:

image

Choosing image and image (positive definite), we get image and so image asymptotically.

The performance of the controller was tested on the WMR of Section 5.7 with the same values of physical parameters and the same desired path using the PD gains given in Table 5.2.

Table 5.2

Gains of the Computed Torque PD Controller

Image

The responses obtained for the image and image trajectories are similar to those shown in Figure 5.17. As an exercise, the reader may test the performance of the controller for a circular desired path with a given radius image (e.g., image). A parameterized form of the circle equation, namely, image, image can be used where image is the angle of the point image with reference to the x-axis of the world coordinate frame. In this case, a suitable polynomial or other series representation of the angle image can be used. A more complex desired path that might be examined is an 8 shaped path of a given size.

References

1. Ogata K. State space analysis of control systems Upper Saddle River, NJ: Prentice Hall; 1997.

2. Asada H, Slotine JJ. Robot analysis and control New York, NY: Wiley; 1986.

3. Spong MW, Vidyasagar M. Robot dynamics and control New York, NY: Wiley; 1989.

4. Wolovich W. Robotics: Basic Analysis and Design. Birmingham, UK: Holt Rinehart and Winston, Dreyden Press;1987.

5. Kanayama Y, Kimura Y, Noguchi T. A stable tracking control method for a nonholonomic mobile robot. IEEE Trans Robot Autom. 1991;7:1236–1241.

6. Yiaoping Y, Yamamoto Y. Dynamic feedback control of vehicles with two steerable wheels. Proceedings of IEEE international conference on robotics and automation. Minneapolis, MN; 1996. 12(1), p. 1006–1010.

7. Panimadai Ramaswamy SA, Balakrishnan SN. Formation control of car-like mobile robots: a Lyapunov function based approach. Proceedings of 2008 American Control Conference. Seattle, Washington; June 11–13, 2008.

8. Tian Y, Sidek N, Sarkar N. Modeling and Control of a nonholonomic wheeled mobile robot with wheel slip dynamics. Proceedings of IEEE symposium on computational intelligence in control and automation. Nashville, TN; March 30–April 2, 2009. p. 7–14.

9. Chang CF, Huang CI, Fu LC. Nonlinear control of a wheeled mobile robot with nonholonomic constraints. Proceedings of 2004 IEEE International conference on systems, man, and cybernetics. The Hague, The Netherlands; October 10–13, 2004. p. 5404–9.

10. Samson C, Ait Abderrahim K. Feedback control of nonholonomic wheeled cart in Cartesian space. Proceedings of IEEE Conference on robotics and automation. Sacramento, CA; 1990. p. 1136–41.

11. Velagic J, Lacevic B, Osmic N. Nonlinear motion control of mobile robot dynamic model. [Chapter 27] In: Jing X-J, ed. Motion Planning. In Tech, Open Books 2008;534–556.

12. Zhang Y, Hong D, Chung JH, Velinsky SA. Dynamic model based robust tracking control of a differentially steered wheeled mobile robot. Proceedings of american control conference (ACC ’88). Philadelphia, PA; June 1988. p. 850–5.

13. Ashoorizad M, Barzamini R, Afshar A, Zouzdani J. Model reference adaptive path following for wheeled mobile robots. Proceedings of international conference on information and automation (IEEE/ICIA’06). Colombo, Sri Lanka; 2006. p. 289–94.

14. Gholipour A, Dehgham SM, Ahmadabadi MN. Lyapunov based tracking control of nonholonomic mobile robot. Proceedings of 10th Iranian conference on electrical engineering. Tabeiz, Iran; 2002. 3, p. 262–69.

15. Aicardi M, Casalino G, Bicchi A, Balestrino A. Closed-loop steering of unicycle-like vehicles via Lyapunov techniques. IEEE Robot Autom Mag. 1995;March:27–35.

16. Lee S, Kim M, Youm Y, Chung W. Control of a car-like mobile robot for parking problem. Proceedings of 1999 IEEE international conference on robotics and automation. Detroit, MI; May 1999. p. 1–5.

17. Lee S, Youm Y, Chung. Control of car-like mobile robots for posture stabilization. Proceedings IEEE/RSJ international conference on intelligent robots and systems (IROS’99). Kyongju, Korea; October 1999. p. 1745–50.

18. Watanabe K, Shiraishi Y, Tang J, Fukuda T, Tzafestas SG. Autonomous control for an omnidirectional mobile robot with feedback control system. [Chapter 13] In: Tzafestas SG, ed. Advances in intelligent autonomous systems. Boston / Dordrecht: Kluwer; 1999;289–308.


1Note that here image is considered to be a column vector, that is, image.

2It is remarked that the error image can also be defined as image. In this case, the feedback gains have opposite signs and actually lead to the same negative feedback controller (Section 5.3.1).

3This is equivalent to considering that our WMR has to track a similar (virtual) differential-drive WMR which is moving with linear velocity image and angular velocity image.

5All the coordinates are those of the center of gravity (and symmetry) Q. The index Q was dropped for notational simplicity.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.138.33.87