Você está na página 1de 6

1580 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO.

9, SEPTEMBER 2002

[7] Wu, Yang, Packard, and Becker, “Induced L -norm control for LPV A more effective control strategy without using the inverse kine-
systems with bounded parameter variation,” Int. J. Robust Nonlinear matics is the task-space control method [1], [11]–[18]. In this method,
Control, vol. 6, no. 9/10, pp. 983–998, 1996. a task oriented information is used directly in the feedback control law.
[8] D. J. Leith and W. E. Leithead, “On formulating nonlinear dynamics
in LPV form,” in Proc. 39th IEEE Conf. Decision and Control, vol. 4, Takegaki and Arimoto [1] proposed a transposed Jacobian controller
Sydney, Australia, 2000, pp. 3526–3527. for setpoint control in Cartesian coordinates. Later, this study is ana-
[9] J. A. Primbs, V. Nevistić, and J. C. Doyle, “On receding Horizon exten- lyzed further by Kelly et al. [11]–[14]. A local feedback control law
sions and control Lyapunov functions,” in Proc. Amer. Control Conf., with imperfect Jacobian matrix from Cartesian space to visual space is
Philadelphia, PA, 1998, pp. 3276–3280.
[10] L. E. Ghaoui and G. Scorletti, “Control of rational systems using linear-
proposed by Miyazaki and Masutani [15]. In these controllers [1]–[15],
fractional representations and linear matrix inequalities,” Automatica, an exact knowledge of the robot kinematics from joint space to task
vol. 32, no. 9, pp. 1273–1284, 1996. space is required. However, since the robot is interacting with its en-
[11] A. Trofino, “Robust stability and domain of attraction of uncertain non- vironment, its kinematics changes for different tasks when it picks up
linear systems,” in Proc. Amer. Control Conf., vol. 4, Chicago, IL, 2000, different objects. To overcome the problem of uncertain kinematics,
pp. 3194–3199.
[12] Y. Huang and W.-M. Lu, “Nonlinear optimal-control: Alternatives to Cheah et al. [16]–[18] proposed task-space feedback laws with uncer-
Hamilton-Jacobi equation,” in Proc. 35th IEEE Conf. Decision and Con- tain kinematics and Jacobian matrix from joint space to task space.
trol, Kobe, Japan, 1996, pp. 3942–3947. In most of the setpoint controllers, an exact knowledge of a gravi-
[13] Y. Huang and A. Jadbabaie, “Nonlinear H control: An enhanced tational force is used in the controllers. When the gravitational force
Quasi-LPV approach,” in Proc. 14th IFAC World Congr., Beijing,
China, 1999, pp. 85–90.
is uncertain, several adaptive control laws [2], [3], [6], [10], [16], [17]
[14] H. K. Khalil, Nonlinear Systems. Upper Saddle River, NJ: Prentice- using a gravity regressor are proposed for compensating the gravita-
Hall, 1996. tional force. However, the exact knowledge of the gravity regressor
[15] M. Mesbahi, M. G. Safanov, and G. P. Papavissilopoulos, “Bilinearity matrix is assumed to be known in these controllers. Unfortunately, no
and complementarity in Robust control,” in Advances in Linear
model can be obtained precisely. In addition, the gravity regressor also
Matrix Inequality Methods in Control, L. E. Ghaoui and S. Niculescu,
Eds. Philadelphia, PA: SIAM, 2000, pp. 269–292. changes when the robot picks up different objects.
[16] E. Feron, P. Apkarian, and P. Gahinet, “Analysis and synthesis of Robust In this note, we propose a task-space adaptive law for setpoint con-
control systems via parameter-dependent Lyapunov functions,” IEEE trol of robot with uncertainties in both the gravity regressor matrix and
Trans. Automat. Contr., vol. 41, pp. 1041–1046, July 1996. kinematics. In addition, we investigate the stability problem when an
estimated task-space velocity is used in the feedback loop. To the best
of our knowledge, such problem has not been studied before. There-
fore, it is unknown whether the stability of the robot’s motion can still
be guaranteed in the presence of such uncertainties. We shall present
sufficient conditions for choosing the feedback gains, gravity regressor,
Task-Space Adaptive Control of Robotic Manipulators and Jacobian matrix to guarantee the stability.
With Uncertainties in Gravity Regressor
Matrix and Kinematics II. PROBLEM FORMULATION
H. Yazarel and C. C. Cheah We consider a class of robotic manipulators with all revolute joints.
These are sometimes said to be articulated robots since their configu-
ration of links and joints corresponds to that of a human arm. In most
Abstract—Thus far, most research in adaptive control of robotic manip- applications, a desired path for the robot end effector is specified in task
ulators has assumed that models of regressor matrix and kinematics are space such as visual space or Cartesian space. Let X 2 Rm represents
known exactly. To overcome these drawbacks, we propose in this note a
task-space adaptive law for setpoint control of robots with uncertainties a task-space vector [16]
in gravity regressor matrix and kinematics. In addition, we investigate the
stability problem when an estimated task-space velocity is used in the feed- X = h(q ) (1)
back loop. Sufficient conditions for choosing the feedback gains, gravity
regressor, and Jacobian matrix are presented to guarantee the stability. where m  n and h(1) 2 Rn ! Rm is generally a nonlinear transfor-
Index Terms—Setpoint control, stability, task space, uncertain kine- mation describing the relation between the joint space and task space.
matics, uncertain regressor. The task-space velocity X_ is related to joint-space velocity q_ as [16]

_ = J (q )q_
X (2)
I. INTRODUCTION
where J (q ) is a Jacobian matrix of the mapping from joint space to
In most applications of robots, a desired path of the end effector is
task space. Note that h(q ) and J (q ) are trigonometric functions of q .
usually specified in task coordinates. However, a majority of the robot
The equation of motion for the robotic manipulator is given in joint
controllers in the literature were joint-space controllers [1]–[10]. In
space as [6]
order to control the robot with these controllers, an inverse kinematics
problem should be solved to generate a desired path in joint coordi- 1 _
nates. M (q )
q+ B0 + M (q ) + S (q; q_) q_ + g (q ) =  (3)
2

where q 2 Rn denotes joint angles, n denotes degrees of freedom


of the robot, M (q ) 2 Rn2n is an inertia matrix, B0 2 Rn2n de-
Manuscript received February 20, 2001; revised February 27, 2001 and April notes a diagonal viscous friction matrix, S (q; q_ )q_ = (1=2)M_ (q )q_ 0
(1=2)f(@=@q )q_ M (q )q_ g , g (q ) = (@P (q )=@q ) 2 R is a gravita-
24, 2002. Recommended by Associate Editor K. M. Grigoriadis. T T T n
tional force,  2 R denotes control inputs, and P (q ) is the potential
The authors are with the School of Electrical and Electronic Engineering, n
Nanyang Technological University, Singapore 639798.
Publisher Item Identifier 10.1109/TAC.2002.802735. energy due to gravitational force. The gravitational force can be com-

0018-9286/02$17.00 © 2002 IEEE


IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO. 9, SEPTEMBER 2002 1581

pletely characterized by a set of parameters ' = ('1 ; . . . ; 'r )T [2],


[3], [6] as

g(q) = Z (q)' (4)

where Z (q ) 2 Rn2r is the gravity regressor.


Several properties of the robot dynamics are given as follows.
Property 1 (Refer to [6] and [8]): The inertia matrix M (q ) is sym-
metric and positive–definite.
Property 2 (Refer to [6] and [8]): The matrix S (q; q_ ) is skew sym-
metric.
Property 3 (Refer to [2], [6], [7], and [8]): For robots with revolute
joints, since g (q ) = (@P (q )=@q )T and P (q ) are trigonometric func-
tions of q , there exists a positive constant kg such that for any constant
(a)

vector qd

kg(q) 0 g(qd )k  kg k4qk (5)


P (q) 0 P (qd ) 0 (q 0 qd )T g(qd ) + 1 kg k4qk2  0 (6)
2

where 4q = q 0 qd .
In most adaptive setpoint controllers proposed in the literature, the
exact gravity regressor matrix Z (1) described in (4) and the exact Ja-
cobian matrix described in (2) are required. In this note, we consider
modeling errors in both the regressor and Jacobian matrices such that
Z (1) and J (1) are uncertain and are estimated by Z^(1) and J^(1) respec-
tively. Hence, the control input is proposed as

 = 0 J^T (q)Kp s(e) 0 J^T (q)Kv X^_ + Z^(q)^ (7)


^_ = 0 LZ^T (q) (q_ + J^+ (q)s(e)) (8)
(b)

where X ^_ = J^(q )q_ , e = X 0 Xd = (e1 ; . . . ; em )T is a positional Fig. 1. (a) Quasinatural potential: S (). (b) Derivative of S ( ): s( ).

deviation from a desired position Xd 2 Rm , Kp and Kv are positive


definite diagonal feedback gains for the position and velocity respec- III. CONTROL WITH UNCERTAIN GRAVITY REGRESSOR
tively, is a positive constant, Z^( q) is an estimated gravity regressor, AND JACOBIAN MATRICES
q 2 Rn is a constant position vector, L 2 Rp2p is a positive–definite Let us define a scalar potential function Si () where  2 R, and its
matrix, si (1), i = 1; . . . ; n are saturated functions to be defined later, derivative si () [6] (see Fig. 1) with the following properties.
the initial estimation of ^ at t = 0 is chosen to be zero without loss of
2) Si () > 0 for  6= 0 and Si (0) = 0.
generality, J^+ (q ) is the pseudoinverse of J^(q ) such that J^(q )J^T (q )
3) Si () is twice continuously differentiable, and the derivative
is nonsingular, and J^(q )J^+ (q ) = I . Clearly, the initial position X (0)
and desired position Xd must also be specified such that J^(q )J^T (q )
si () = (dSi ()=d) is strictly increasing in  for jj < i
with some i and saturated for jj  i , i.e., si () = 6si
for   + i and   0 i respectively where si is a positive
is nonsingular at these positions. Note that in the controller (7), the
task-space vector X is measured by a sensor such as vision systems,
constant.
electromagnetic measurement systems, and laser tracking systems, and
^_ is estimated as J^(q )q_ . 4) There are constants ci > 0, di > 0, di (> di ) > 0 such that
the velocity vector X
We assume that the approximate Jacobian matrix J^T (q ) is chosen
di si2 () si ()  di si2 () > 0
Si () ci si2 () for  6= 0:
so that
(10)
kJ T (q) 0 kp
J^T (q) (9)
Some examples of saturation function can be found in [6] and [8]. If
where p is a positive constant to be defined later. Note that J^(q ) is not we substitute (4) and (7) into (3), we have
required to have the same structure as J (q ) so long as the error of the
approximate Jacobian matrix in (9) is bounded. M (q)q + B0 + 1 M_ (q) + S (q; q_) q_ + g(q) 0 g(qd )
2
Remark 1: In [16] and [17], the actual task-space velocity X_ and
exact gravity regressor matrix are used in the controllers. Since X is ^_ + J^T (q )Kp s(e) + Z (qd )' 0 Z^(
+J^T (q )Kv X q)^ = 0 (11)
measured by a sensor, it is usually noisy to obtain X_ by differentiating
X . In this note, we estimate the task velocity from joint velocity, which where g (qd ) = Z (qd )'. Since manipulator has its own configuration at
is less noisy. The main difficulty of estimating the task velocity from the joint space, we define qd as a desired joint position such that h(qd ) =
joint velocity comes from the fact that the Jacobian matrix is uncertain Xd . However, note that qd , g(qd ), and Z (qd ) are introduced for the
in the presence of kinematic uncertainties. As a result, it is not sure purpose of analysis and are not required in the controller as seen from
whether the stability of the system could still be guaranteed. (7) and (8).
1582 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO. 9, SEPTEMBER 2002

We define a Lyapunov-like function V as =


By choosing L lI where l is a positive scalar, a sufficient condition

= 21 q_T M (q)q_ + q_T M (q)J^+ (q)s(e)


to guarantee (20) is
V
m lmin Z^(q)Z^T (q) 0 max Z (qd)Z T (qd)  0 (21)
+ (kpi + kvi )Si (ei )
i=1 where

+P (q) 0 P (qd ) 0 1qT g(qd ) + 21 (' + w)T (' + w) min Z^(q)Z^T (q) > 0: (22)

0 12 wT w + 21 ^T L01^ (12) Therefore, from (17), (19), and (21)

 41 q_T M (q)q_ +
m
where V f (kvi ci 0 m )g si2 (ei )
t i=1
Z T (qd )y( )d
+ kks(e)k + 1 (' + w)T (' + w)
w (13) 2
0
2
1q = q 0 qd , y = q_ + J^ (q)s(e), kpi , kvi denote the ith diagonal
+ + lmin [Z (q)Z^T (q)]
^
elements of Kp and Kv , respectively. Since
0max [Z (qd)Z T (qd )] kY k2  0 (23)
1 q_T M (q)_q + q_T M (q)J^+ (q)s(e) + m kvi Si (ei )
4 i=1
and, hence, V is a nonnegative function.
1 Differentiating V in (12) with respect to time and using property 1),
^
= 4 (q_ + 2 J (q)s(e)) M (q)(_q + 2 J^+ (q)s(e))
+ T
we have

=q_T M (q)q + 12 q_T M_ (q)_q + s(e)T (J^+ (q))T M (q)q


m
0 2 s(e)T (J^+ (q))T M (q)J^+ (q)s(e) + kvi Si (ei ) V_
m
i=1
+ s(e)T (J^+ (q))T M_ (q)_q
 f (kvi ci 0 m )gsi (ei )2
T (J^+ (q ))T M (q )q_ + s(e)T J_^+ (q )
T
i=1
(14)
+ _
s (e ) M (q)q_
where m max [(J^+ (q))T M (q)J^+ (q)] and max [A] denotes the + X_ (kp + kv )s(e)
T

maximum eigenvalue of a matrix A. From property 3), since g (q ) 0 + q_T (g(q) 0 g(qd )) + 'T w_ + ^T L01^_ (24)
g(qd ) satisfies (5) and P (q) 0 P (qd ) 0 1q g(qd ) satisfy (6), and
T
since J^+ (q ) exists and is bounded, Kp can be chosen large enough so where w_ = Z T (qd ) q_ + J^+ (q )s(e) . Next, substituting M (q ) q
that the following conditions are satisfied for a sufficiently small k [6]: _
^
from (11) and  from (8) into (24) and using property 2), we have
n
P (q) 0 P (qd ) 0 1qT g(qd ) + kpi Si (ei ) V_ = 0W (25)
i=1
 kks(e)k2 (15) where
s(e) (J^+ (q))T fg(q) 0 g(qd )g + s(e)T Kp s(e)
T
W = sT (e)(J + (q))T (g(q) 0 g(qd )) + sT (e)Kp s(e)
 kks(e)k2: (16)
+ q_T J^T (q)Kv J^(q) + B0 q_
Note also that each component si (ei ) satisfies (10) and Si (ei ) is
quadratic in the vicinity of e = 0 (see Fig. 1.). If we substitute (14)
0 q_T J T (q) 0 J^T (q) Kp s(e)
and (15) into (12) 0 sT (e)Kv (J (q) 0 J^(q))_q
 14 q_T M (q)_q + B0 0 1 M_ (q) + S (q; q_) q_
m
V f (kvi ci 0 m )gs2i (ei ) + kks(e)k2 + s(e)T (J^+ (q))T 2
i=1
+ 21 (' + ) ( + w) 0 21 wT w + 21 ^T L01^
wT ' (17) 0 s_ (e)T (J^+ (q))T M (q)_q
0 s(e)T J_^ (q) M (q)q_
+ T
where kvi can be chosen large enough to satisfy the inequality : (26)

kvi ci 0 m > 0: (18)


()
From the last term of (26), since s e is bounded, there exist con-
( ) ^()
Since Z qd and Z q are constant, from (13) and (8), the last two 0 0
stants c0 > and c1 > so that [6] (27), as shown at the bottom of
terms of (17) can be expressed as the next page, holds true. Let 1 () ^()
J T q 0 J T q . Substituting (16)
1 ^T L01^ 0 1 wT w = 1 Y (t)T (Z^(q)LZ^T (q) and (27) into (26), we have
2 2 2 W q_T (J^T (q)Kv J^T (q) + B0 0 c0 I )q_
0Z (qd )Z T (qd))Y (t) (19) + (k 0 c1 )s(e)T s(e) 0 q_T 1(Kp + Kv )s(e)
where Y (t) = 0 y ( )d and ^(0) = 0. Hence, (19) is nonnegative if
t  min J^T (q)Kv J^(q) + B0 0 c0 kq_k2
0 p(max [Kp ] + max [Kv ])ks(e)k 1 kq_k
min Z^(q)LZ^T (q) 0 Z (qd )Z T (qd )  0: (20) + (k 0 c1 )ks(e)k2: (28)
IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO. 9, SEPTEMBER 2002 1583

Since

0ks(e)k 1 kq_k  0 12 (ks(e)k2 + kq_k2 ) (29)

we have

W  (max [Kv ]la 0 c0 )kq_k2


+(max [Kv ]lb 0 c1 )ks(e)k2 (30)

where

la =1 0 p (a + ) lb = a k 0 p (a + )
2 max [Kp ] 2
 J T (q)Kv J (q) + B0
1 = min a = max [Kp ] :
max [Kv ] max [Kv ]
Hence, if
2a  k
21 [K ]
min ; >p (31)
a+ a+
then la > 0 and lb > 0 and Kv can be chosen large enough so that Fig. 2. Variation of p with a.

la 0 c0 > 0 lb 0 c1 > 0: (32)


max [Kv ] max [Kv ] the theorem are simple conditions to achieve in practice. Conditions
(15), (16), (18), and (32) simply mean that the feedback gains should be
Therefore, W is positive–definite in q_ and s(e). chosen sufficiently large. Conditions (9) and (31) imply that the feed-
A graphical illustration of (31) is shown in Fig. 2. As seen from this back gain ratio a = kp =kv should be small to allow a larger range of
figure, the allowable bound p of Jacobian uncertainty is inversely pro- uncertainty in the Jacobian matrix. Finally, (20) or (21) simply mean
portional to the ratio a of positional feedback gain to velocity feedback that the gain L should be chosen sufficiently large. Hence, tuning can
gain. If the uncertainty p increase, a smaller a is required. Hence, to be established easily in practice.
allow more uncertainty in Jacobian matrix, a should be kept smaller. Remark 4: Note that in the absent of the saturation function, the
As compared with the stability condition in our previous papers [16], theory is still valid in a local sense. In addition, since the task-space
[17], we note from (31) that a could not be too small with the estimated positional error is usually bounded because the robot workspace is fi-
task space velocity. This could be due to the fact that additional uncer- nite, the saturation function may not be needed for most applications.
tainty is present in the task velocity (estimated by J^(q )q_ ) and, hence,
Kv could not be too large. IV. SIMULATION RESULTS
We are now in a position to state the following.
Theorem: The closed-loop system with uncertain Jacobian matrix Let us a consider a three-link planar robot manipulator as shown in
J^(q) and with uncertain gravity regressor Z^(1), described by (8) and Fig. 3. The gravitational regressor matrix of the robot can be repre-
(11) gives rise to asymptotic convergence of (X; q_ ) to (Xd ; 0) as t ! sented as [20]
1, if feedback gains Kp and Kv are chosen to satisfy (15), (16), (18), c1 c12 c123
(31), and (32), J^(q ) is chosen to satisfy the condition (9) and Z^(1) and Z (q) =
0 c12 c123 (33)
L are chosen to satisfy (20) or (21) and (22). 0 0 c123
Proof: Since W is positive–semidefinite, from (25), we have
where c1 = cos(q1 ), c12 = cos(q1 +q2 ), and c123 = cos(q1 +q2 +q3 ).
V_ = 0W  0: If a fixed camera is placed some distance away from the robot, the
Jacobian matrix J (q ) of the mapping from joint space to visual space
Hence, V is a nonnegative Lyapunov-like function whose time deriva- is given by [18], [20]
tive is negative in (s(e); q_). Since W = 0 means s(e) = 0 and q_ = 0,
by LaSalle’s invariance theorem, we have e ! 0, q_ ! 0, as t ! 1. J (q) = f1
cos  sin 
Remark 2: The result is valid globally in a workspace such that z 0 f1 0 sin  cos 
J^(q)J^T (q) is nonsingular. Hence, if the estimated Jacobian matrix
J^(q) is designed to be of full rank at the singular points, it is possible 2 0ll11sc11 0+ ll22sc1212 0+ ll33sc123
123 0l2 s12 0 l3 s123 0 l3 s123
l2 c12 + l3 c123 l3 c123
to overcome the singularity problem [19]. (34)
Remark 3: The results in this note are qualitative results to guar-
antee the stability of the robot in presence of uncertain kinematics and where s1 = sin(q1 ), s12 = sin(q1 + q2 ), and s123 = sin(q1 + q2 + q3 ),
gravity regressor matrix. In the presence of kinematic uncertainty, a  represents the angle of rotation of the vision coordinates relative
joint-space controller [20] cannot be used. The stability conditions in to Cartesian coordinates, the offset of the origins of the coordinates

T + T
s(e)T J^+ (q) B0 0 1 M_ (q) + S (q; q_) q_ 0 s_ (e)T (J^+ (q))T M (q)q_ 0 s(e)T J_^ (q) M (q)q_
2

 0 c0 kq_k2 0 c1 ks(e)k2 (27)


1584 IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO. 9, SEPTEMBER 2002

(a)

Fig. 3. A three-link planar robot.

d = (dx ; dy )T (see Fig. 3) were set as 0 m, z is the perpendicular dis-


tance between the robot and the camera, f1 is the focal length of the
camera, and li is the length of link i. In the simulations, l1 , l2 , and l3
were set as 0.5 m, f1 was chosen as 0.05 m, z was chosen as 3 m, 
was chosen as zero, the masses of all links m1 , m2 , m3 were chosen as
1 kg, the viscous friction of coefficient of all joints was set as 25, and
g = 9:8 m/s2 . The mass and length of the object were set as 0.5 kg and
0.2 m, respectively. The robot was required to move from the initial
position [q1 (0); q2 (0); q3 (0)]T = [0; 0; 0=18] or [x(0); y (0)]T =
[1:492; 00:1736] m to the desired position [xd ; yd ]T = [1:25; 0:25] m.
The constant was chosen as 0.3. ^(0) was set as zero without loss of (b)
generality. This setting is required in the theoretical proof [see (19)],
Fig. 4. (a) Stable response for the first task. (b) Stable response for the second
and it means that the parameters is completely unknown. In the pres- task.
ence of uncertainty in camera parameters and in kinematics, an approx-
imate Jacobian matrix is obtained as shown in
where c123 = cos(q1 + q2 + q3 0 ) and is the angle between
f^1 cos ^ sin ^
J^(q) = the third link of the robot and the object. Using our results, switching
z^ 0 f^1 0 sin ^ cos ^ of controllers with different regressor and Jacobian matrices are not

2 0^l1l1cs11+0^l2l2cs1212+0^l3l3cs123123 0^l2 s12 0 ^l3 s123 0^l3 s123


^ ^ ^ required as the changes can be treated as a modeling error. As a result,
^l2 c12 + ^l3 c123 ^l3 c123 the robot system becomes more flexible to deal with unforseen changes
(35) in its dynamics and kinematics. Note that, although the dimension of
the regressor matrix has changed, sufficient condition (21) can still be
where f^1 , z^, ^, ^
l1 , ^l2 , and ^l3 are the estimations of f1 , z ,  , l1 , l2 , and satisfied. In the simulation, the regressor matrix was chosen as
l3 , respectively. For the simulations, we set f^1 = 0:1 m, z^ = 2 m,
^ = 30 , ^l1 = 0:75 m, ^l2 = 0:60 m, and ^l3 = 0:25 m. c1 c12 c123
In this note, we consider an example where the dimension of the Z^(q) = 0 c12 c123 (36)
gravity regressor is modeled wrongly due to changes in tasks. When 0 0 c123
the robot picks up an object, the actual gravity regressor changes from
(33) to and the feedback control gains were chosen as Kp = 500I , Kv =
15 000I and L = 200I . q1 , q2 and q3 were chosen arbitrary as =8,
c1 c12 c123 c123 =6 and =3, respectively. Fig. 4(a) shows the position errors of the end
Z (q) = 0 c12 c123 c123 effector when the robot does not pick up an object and Fig. 4(b) shows
0 0 c123 c123 the position errors when the robot picks up the object with = =6.
IEEE TRANSACTIONS ON AUTOMATIC CONTROL, VOL. 47, NO. 9, SEPTEMBER 2002 1585

Both results show stable responses using the same regressor matrix [13] , “Fixed camera visual servo control for planar robots,” in Proc.
(36). IEEE Int. Conf. Robotics and Automation, Minneapolis, MN, 1996, pp.
Remark 5: As seen from (7) and (35), the feedback gains Kp in the 2643–2649.
(^ ^ ^) =
simulation was multiplied by a smaller scaling factor f1 =z 0 f1
[14] R. Kelly, R. Carelli, O. Nasisi, B. Kuchen, and F. Reyes, “Stable vi-

(0 1 2 0 1) = 0 05263
: = 0 :
sual servoing of camera-in-hand robotic systems,” IEEE/ASME Trans.
: and Kv was multiplied by the square of the Mechatron., vol. 5, pp. 39–48, Mar. 2000.
factor. Hence, the overall gains for the position and velocity were not [15] F. Miyazaki and Y. Masutani, “Robustness of sensory feedback control
( ^ ^ ^ ) = 0 05263 500 = 26 31
high and were given as f1 =z 0 f1 Kp : 2 I : based on imperfect Jacobian,” in Robotic Research: Fifth Int. Symp.,
^ ^
( ^ ) = (0 05263) 15000 = 41 55
and f1 =z 0 f1 2 Kv : 2
2 I : , respectively.
[16]
1990, pp. 201–208.
C. C. Cheah, S. Kawamura, and S. Arimoto, “Feedback control for
robotic manipulators with an uncertain Jacobian matrix,” J. Robot.
Syst., vol. 12, no. 2, pp. 119–134, 1999.
[17] , “Feedback control for robotic manipulators with uncertain kine-
V. CONCLUSION matics and dynamics,” in Proc. IEEE Int. Conf. Robotics and Automa-
tion, Leuven, Belgium, 1998, pp. 3607–3612.
[18] C. C. Cheah, K. Lee, S. Kawamura, and S. Arimoto, “Asymptotic sta-
In this note, we have proposed an adaptive task-space controller for bility of robot control with approximate Jacobian matrix and its applica-
setpoint control of robotic manipulator with uncertain gravity regressor tion to visual servoing,” in Proc. IEEE Int. Conf. Decision and Control,
matrix and uncertain Jacobian matrix. The main advantage of the pro- Sydney, Australia, 2000, pp. 3939–3944.
posed adaptive law is that exact knowledge of the gravity regressor and [19] C. C. Cheah, K. Li, S. Kawamura, and S. Arimoto, “Approximate Ja-
cobian feedback control of robotic manipulators and its passability at
kinematics is not required. Another advantage is that the task-space singular points,” in Proc. Int. Conf. Control, Automation, Robotics, Vi-
damping is estimated from joint velocity and, hence, is less noisy. We sion, Singapore, 2000.
have shown that asymptotic convergence can be guaranteed even with [20] H. Yazarel, C. C. Cheah, and H. C. Liaw, “Adaptive SP-D control of
uncertain gravity regressor matrix and uncertain Jacobian matrix. Suf- robotic manipulator in presence of modeling error in gravity regressor
ficient conditions for choosing the feedback gains and gravity regressor matrix: Theory and experiment,” IEEE Trans. Robot. Automat., vol. 18,
pp. 373–379, June 2002.
matrix are presented to guarantee the stability. Simulation results illus-
trate the performance of the proposed controller.

ACKNOWLEDGMENT
Correction to “Finite Horizon State-Feedback Control
The authors would like to thank the reviewers for their constructive of Continuous-Time Systems With State Delays”
suggestions.
E. Fridman and U. Shaked

REFERENCES In the above paper,1 a typographical error appeared that needs to be


corrected. The second inequality of (18) should read
[1] M. Takegaki and S. Arimoto, “A new feedback method for dynamic con-
0 0 0
trol of manipulators,” ASME J. Dyn. Syst., Meas., Control, vol. 102, pp.
119–125, 1981. 2 0(
y s R) (0; s; )y()dds  h 0(
y s )12 y(s)ds
[2] P. Tomei, “Adaptive pd controller for robot manipulators,” IEEE Trans. 0h 0h 0h
0 0
Robot. Automat., vol. 7, pp. 565–570, Aug. 1991.
+ 0(
y s ) R (0; s; )1201R(0; ; s)dy(s)ds:
[3] R. Kelly, “Comments on adaptive pd controller for robot manipulators,”
IEEE Trans. Robot. Automat., vol. 9, pp. 117–119, Jan. 1993. 0h 0h
[4] J. Wen, K. Kreutz-Delgado, and D. Bayard, “Lyapunov function-based
control laws for revolute robot arms,” IEEE Trans. Automat. Contr., vol. The correct version of the latter inequality has been applied to derive
37, pp. 231–237, Feb. 1992. (16).
[5] S. Arimoto, “Fundamental problems of robot control: Part I, innovations
in the realm of robot servo-loops,” Robotica, vol. 13, pp. 19–27, 1995.
[6] , Control Theory of Nonlinear Mechanical Systems—A Pas- ACKNOWLEDGMENT
sivity-Based and Circuit-Theoretic Approach. Oxford, U.K.:
Clarendon, 1996. The authors would like to thank W.-S. Yu of the Department of Elec-
[7] R. Kelly, “Pd control with desired gravity compensation of robotic ma- trical Engineering, Tatung University, Taiwan, for drawing their atten-
nipulators: A review,” Int. J. Robot. Res., vol. 16, no. 5, pp. 660–672, tion to the error.
1997.
[8] , “Global positioning of robot manipulator via pd control plus a
class of nonlinear integral actions,” IEEE Trans. Automat. Contr., vol.
43, pp. 934–938, July 1998.
[9] J. J. E. Slotine and W. Li, “Adaptive manipulator control: A case study,”
IEEE Trans. Automat. Contr., vol. 33, pp. 995–1003, Nov. 1988.
[10] A. Laib, “Adaptive output regulation of robot manipulators under actu-
ator constraints,” IEEE Trans. Robot. Automat., vol. 16, pp. 29–35, Jan. Manuscript received April 4, 2002. Recommended by Associate Editor Y.
2000. Ohta.
[11] R. Kelly and A. Coello, “Analysis and experimentation of transpose The authors are with the Department of Electrical Engineering-Systems, Tel
Jacobian-based cartesian regulators,” Robotica, vol. 17, pp. 303–312, Aviv University, Tel Aviv 69978, Israel (e-mail: emilia@eng.tau.ac.il).
1999. Publisher Item Identifier 10.1109/TAC.2002.802751.
[12] R. Kelly, “Regulation of manipulators in generic task space: An en-
ergy shaping plus damping injection approach,” IEEE Trans. Robot. Au- 1E. Fridman and U. Shaked, IEEE Trans. Automat. Contr., vol. 45, pp.
tomat., vol. 15, pp. 381–386, May 1999. 2406–2411, Dec. 2000.

0018-9286/02$17.00 © 2002 IEEE

Você também pode gostar