Sunteți pe pagina 1din 16

Maximum-dynamic-payload trajectory for exible robot

manipulators with kinematic redundancy


Shigang Yue
a,*
, S.K. Tso
a
, W.L. Xu
b
a
Department of Manufacturing Engineering and Engineering Management, City University of Hong Kong, Kowloon,
Hong Kong
b
Institute of Technology and Engineering, Massey University, Palmerston North, New Zealand
Received 8 December 1999; received in revised form 28 November 2000; accepted 8 December 2000
Abstract
High payload-to-mass ratio is one of the advantages of exible robot manipulators. For a given point-to-
point task, the maximum dynamic payload-carrying capacity of a exible redundant robot manipulator
(FRM) is established in the paper. A nite-element model is presented for describing the dynamics of the
system considering the added payload and the exibility of the FRM. Dierent from rigid-body manipu-
lators, vibration deformation, which is due to lightweight and high operation accelerations, is considered as
constraint for the system optimization. The kinematic redundancy is integrated into the optimization
formulation. The maximum-dynamic-payload trajectory is determined through optimization in which a
planar FRM is employed as an example. This method is useful especially for FRMs performing repeated
point-to-point operations. 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Redundancy; Flexible robot manipulators; Dynamic payload; Trajectory; Optimization
1. Introduction
The dynamic payload-carrying capacity of a robot manipulator is usually dened as the
maximum payload that the manipulator can repeatedly lift in its fully extended conguration
while the dynamics of both the load and the robot manipulator itself must be taken into account.
Given two endpoints of the robot manipulator, the problem is to synthesize a dynamic trajectory
for the end-eector that would allow carrying the maximum load between the two points. For the
rigid manipulators, the major limiting factor in determining the maximum load is the joint ac-
tuator capacity. The exible deformation, which results from the lightweight structure and high
Mechanism and Machine Theory 36 (2001) 785800
www.elsevier.com/locate/mechmt
*
Corresponding author.
E-mail address: shigang@informatik.uni-kl.de (S. Yue).
0094-114X/01/$ - see front matter 2001 Elsevier Science Ltd. All rights reserved.
PII: S0094- 114X( 00) 00059- 8
operating accelerations, indicates that precise tracking of the synthesized trajectory is dicult.
Deformation constraint should also be considered in the trajectory synthesis for precision end-
point tracking of exible robot manipulators.
Wang and Ravani [9,10] studied the problem associated with the load-carrying capacity of
mechanical manipulators, and found that the problem of synthesizing a point-to-point dynamic
robot motion with optimum load-carrying capacity can be formulated as a trajectory optimization
problem. They solved the optimal trajectory problems by using iterative linear programming
(ILP). However, this method is feasible only when the manipulator arm is considered to be rigid.
When the manipulator is composed of lightweight links and operates at high speeds, the torques
derived for a corresponding rigid manipulator will not drive the exible manipulator exactly along
the desired end-eector trajectory.
The use of lightweight robotic manipulators has been recognized as a possible solution to in-
crease the load-to-mass ratio and the operation speed. Other potential advantages of exible links
include lower energy consumption, use of smaller actuators, safer operation due to reduced in-
ertia, easier transport, a more compliant structure for assembly and possible elimination of
gearing. Much work done in the past decades on exible robot manipulators involves mainly
modeling, vibration control and trajectory tracking control as reviewed and discussed in [18].
Regarding the dynamic payload capacity of exible manipulators, Tu and Rastegar [11] studied
the eects of payload on the vibrational excitation of a 2R robot manipulators during motion.
Parks and Pak [12] studied the eects of payload on the dynamics of a exible manipulator
through a single-link arm. Korayen and Basu [13] presented a formulation and numerical solution
for the problem of nding a point-to-point trajectory with maximum load-carrying capacity, and
an example with a 2R exible manipulator was also given. However, the above research focused
on exible manipulators with only one or two links and little has been done on the dynamic
payload capacity of exible robot manipulators with multiple links, especially those with kine-
matic redundancy.
Recently, researchers have found that redundancy, which is widely used in robot manipulators
to provide extra performance while tracing a given end-eector trajectory [14,15], may have a
special use for vibration reduction of multiple-link exible robot [1619]. In fact, the redundancy
also provides the possibility for an exible redundant robot manipulator (FRM) to increase its
dynamic payload-carrying capacity.
In the following sections, the dynamic payload capacity of FRM is studied. First, a nite-el-
ement model is modied and used to describe the dynamic behavior of multiple-link exible robot
manipulators with payload. For a given point-to-point task, the problem of nding the optimal
trajectory for maximum dynamic payload which meets the constraints of the maximum vibration
deection and joint torques is formulated. In the last part of this paper, a planar FRM is used as
an example in a numerical simulation study and the results obtained are presented.
2. Dynamic model formulation
A nite-element method is used here to derive the dynamic control equations of robot ma-
nipulators with multiple exible links. When only link exibility is considered, the deformation of
links can be described by a basic element [20].
786 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
The transverse deections of a basic element are modeled by a quintic polynomial and the
longitudinal deections assumed to be a linear polynomial. The generalized coordinates of the
basic element are assembled in a matrix form:
/ = /
1
/
2
/
8

T
; (1)
where /
1
and /
5
are the axial displacements along the x-axis, /
2
and /
6
are the transverse dis-
placements along the y-axis, /
3
and /
7
are the rotary displacements about the z-axis, and /
4
and
/
8
are the curvature displacements.
Dynamic control equations of the element can be obtained as
[m
e
[

/ [c
e
[
_
/ [k
e
[/ = p
e
q
e
f
e
; (2)
where [m
e
[ is the element mass matrix, [c
e
[ is the element damping matrix, [k
e
[ is the element
stiness matrix, p
e
is the element inertial force matrix, q
e
is the element external force matrix,
and f
e
is the matrix of external force from the adjacent elements.
The exible deformation of the robot system can be described by global generalized coordinates
U = U
1
U
2
U
i
U
nu

T
; (3)
where nu denotes the total number of the generalized coordinates, U
i
represents the ith system
generalized coordinates. The dynamic equations of a exible robot manipulator system can be
obtained by assembling the dynamic equations of all the elements [20] together:
[M[

U [C[
_
U [K[U = P; (4)
s = ([D[ [J
z
[)q H E; (5)
where [M[ is the nu nu global mass matrix, [C[ is the nu nu global damping matrix, [K[ is the
nu nu global stiness matrix, P is the nu 1 inertial-force matrix, U,
_
U and

U are the
generalized coordinates, generalized velocity vector and generalized acceleration vector, respec-
tively, which describe the deformation behavior of the exible links; [D[ is the n n inertia mass
matrix, q and q are the vectors which describe the joint angle and angular acceleration re-
spectively, H is the n 1 centrifugal, coriolis and gravitational terms of rigid and exible
couplings, E is the n 1 link-exibility term, [J
z
[ is the n 1 rotor inertia-mass matrix, s is the
n 1 actuator-torque matrix, and n is the joint number of the robot system. The above matrices
are in general functions of q; _ q and q.
When a payload is carried by the end-eector, it can be assumed that the payload is rigidly
grasped by the hand. This means that the load and the end-eector can be treated as a single
composite rigid body. The center of mass of this composite body is denoted by c as shown in
Fig. 1, and the total mass of this composite body m
c
is,
m
c
= m
e
m
p
; (6)
where m
e
is the mass of the end eector, and m
p
is the mass of the payload. The moment of inertia
of the composite body about the original point o of the (n 1)th coordinate system (X
n1
oY
n1
) is
denoted by I
o
and is given by
I
o
= I
e
m
e
l
2
e
I
p
m
p
l
2
p
; (7)
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 787
where I
e
is the moment of inertia of the end-eector about its center of mass e, I
p
is the moment of
inertia of the payload about its center of mass p, l
e
is the distance from the origin of the (n 1)th
coordinate system to the center of mass of the hand, l
p
is the distance from the origin of the
(n 1)th coordinate system to the center of mass of the payload.
To incorporate the added payload mass at the end-eector, the mass matrix of the above dy-
namic control equation should be modied as
[M[
p
= [M[ [DM[; (8)
where [M[
p
is the nu nu modied global mass matrix; [DM[ is the added mass matrix due to the
payload. [DM[ can be obtained by adding mass and inertia into the mass matrix through
[DM[ = [T
a
[
[m
e
[
p
0
_ _
T
(9)
where [T
a
[ is the assembled transform matrix [20,21], and [m
e
[
p
is the matrix which contains the
payload mass and inertia:
[m
e
[
p
=
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 I
o
0
0 0 0 0 0 0 0 0
0 0 0 0 m
c
0 0 0
0 0 0 0 0 m
c
0 0
0 0 I
o
0 0 0 0 0
0 0 0 0 0 0 0 0
_

_
_

_
(10)
Since coupling terms exist in damping, stiness and force matrices, these matrices should be
modied respectively in the same way. The modied global damping matrix [C[
p
can be obtained
as
[C[
p
= [C[ [T
a
[
[c
e
[
p
0
_ _
T
; (11)
Fig. 1. End-eector and payload as a composite rigid body.
788 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
where [c
e
[
p
is given as
[c
e
[
p
=
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 2m
c
x 0 0
0 0 0 0 2m
c
x 0 0 0
0 0 0 0 0 0 0 0
_

_
_

_
; (12)
where x is the angular velocity of the composite body. The modied global stiness matrix [K[
p
can be obtained as
[K[
p
= [K[ [T
a
[
[k
e
[
p
0
_ _
; (13)
where [k
e
[
p
is given as
[k
e
[
p
=
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
0 0 0 0 m
c
x
2
m
c
e 0 0
0 0 0 0 m
c
e m
c
x
2
0 0
0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0
_

_
_

_
; (14)
where e is the angular acceleration of the composite body. The modied force matrix P
p
can be
obtained as
P
p
= P [T
a
[
p
e

p
0
_ _
T
; (15)
where p
e

p
is given as
p
e

p
=
0
0
I
o
e
0
(x
2
l a
x
)m
c
(a
y
el)m
c
0
0
_

_
_

_
; (16)
where a
x
and a
y
are the accelerations of the composite body in the x-axis and y-axis, respectively,
and l is the length of the end element.
Thus the modied dynamic equation of the FRM system, which describes the dynamic be-
havior of the payload and the robot at the same time, can be written as
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 789
[M[
p


U [C[
p

_
U [K[
p
U = P
p
: (17)
Flexible deformation of a FRM system can be obtained by solving the above equations if the
nominal motion prole is given. Therefore the computed torque s of the exible robot can be
acquired from Eq. (5), in which exible vibration is taken into account through the exibility term
E.
3. Joint motion planning and initial congurations of FRM
In the present study we assume that the number of rigid degrees of freedom n is greater than the
number of end-eector degrees of freedom m. This means that, for a given position/posture of the
end-eector, there is an innite number of ways to move the joints to that position. The self-
motion capacity allows the links to move while maintaining the end-eector in a xed point, which
has been the subject of much research investigation. For an FRM, the innite number of ways to
move its joints and links while keeping its end-eector xed provides the possibility to improve its
payload capacity and other dynamic behaviors.
3.1. Joint motion planning of a redundant robot
The real position of a exible robot's end-eector is a function of not only the joint angle vector
q, but also the exible link deformations [20]. However, at the stage of joint-motion planning of
an FRM, only nominal joint motion is considered.
The nominal end-eector position of an FRM can be determined by the joint-angle vector q
if all the links are considered as rigid bodies:
x = f (q); (18)
where x is the set of coordinates of the end-eector, x R
m
.
Based on the above equation, we nally obtain [14,15]:
q = [J

r
[(x [J
r
[_ q) ([I[ [J

r
[[J
r
[)e; (19)
where [J
r
[ is the m n rigid Jacobian matrix, [J[ R
mn
is the pseudoinverse of the rigid Jacobian
matrix [J
r
[ , and [J

r
[ can be obtained as
[J

r
[ = [J
r
[
T
([J
r
[[J
r
[
T
)
1
: (20)
x is the acceleration of the end-eector, [I[ R
nn
is a unit matrix, e R
n
is a vector of null
space of the redundant robot, and ([I[ [J

r
[[J
r
[)e N(J) is the homogeneous solution. The
homogeneous solution is the self-motion between the links of the redundant robot manipulator
that results in no end-eector motion.
If the arbitrary vector e is set to be zero (i.e., e = 0), we also have the following joint
planning equation:
q = [J

r
[(x [J
r
[_ q): (21)
The joint motion planned by the above equation has minimum joint angular acceleration [14].
Less vibration deformation may result if the joint motion of an FRM is planned according to the
790 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
above equation because joint angular acceleration has been found to be one of the main exciting
sources of vibration [17].
3.2. Initial conguration of FRM
For a redundant robot manipulator, an innite number of possible congurations may be
posed within the self-motion space. Therefore, which starting conguration should the FRM
assume is one of the problems to be solved by means of an optimization process.
The nominal motion giving the tentative congurations of the FRM can be described by the
following equations with the end-eector's velocity and acceleration maintained zero:
q
0
= [J

r
[(0 [
_
J
r
[_ q) ([I[ [J

r
[[J
r
[)e
0
; (22)
where zero trajectory velocity _ x and acceleration x imply no nominal end-eector motion,
that is, the nominal endpoint remains xed. The second part of the above equation describes the
self-motion between the innite tentative congurations.
The tentative congurations can be represented by the joint-angle vector q
0
and in fact it is a
function of endpoint position and arbitrary vector e:
q
0
= f (x
i
; y
i
; e); (23)
where (x
i
; y
i
) is the point of the tentative trajectory.
For a point-to-point trajectory (especially for a repeated trajectory), the above equations
mean that, at the start, the tentative conguration can be altered to a suitable one in order to
gain optimal dynamic behaviors. In the following formulation of optimization for maximum
payload, the tentative conguration of the FRM is employed as one of the possible congu-
rations.
4. Formulation of the optimal trajectory problem
The problem of synthesizing the dynamic robot trajectory with maximum payload can be
formulated as a trajectory optimization problem based on the modied Eqs. (5) and (17). By
considering point-to-point motions with actuator, joint-variable and vibration-deection con-
straints throughout the trajectory, the formulation can be written as one of maximization
Max : m
p
: (24)
During the motion, the vibration deection of the FRM may cause serious tracking errors.
Therefore, vibration deformation constraints must be considered in the process of optimizing the
FRM trajectory for maximum payload capacity in order to keep the tracking errors at acceptable
levels,
d
a
6d 6d
a
; (25)
where d is the endpoint deformation vector, and d
a
is the allowed endpoint deformation
vector.
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 791
Currently the most popular choice of actuators for lightweight manipulators is the permanent-
magnet DC motor. The torque of these motors can be constrained according to:
s

a
6s 6s

a
; (26)
where s

a
and s

a
are the upper and lower allowed torque vectors. The relationship between
the allowed torque and angular speed of these motors may be described by the following linear
equations [9,13]:
s

a
= s
0
k_ q; (27)
s

a
= s
0
k_ q; (28)
where s
0
is the nominal allowed torque, _ q is the joint angular velocity vector and k is the
coecient.
Since there is structural limitation at every joint, the joint angle is bounded by
q
l
6q 6q
u
; (29)
where q is the joint angle vector, and q
l
and q
u
are the lower and upper bounds of the joint
angle vector.
The trajectory-synthesis problem formulated above is a nonlinear optimization problem.
Dierent from a general optimal control problem where the objective function has an integral
form, the above objective function consists of a variable m
p
which is not a function of time.
The synthesized point-to-point trajectories should be discretized into w points during the
process and constraints should be satised at every point. Therefore, the maximum payload
for a particular trajectory is the minimum one among the maximum payloads over the w
points.
5. Trajectory proles
Suppose that movements of the robot are constrained to one plane (such as horizontal or
vertical plane) and the end-eector moves along a point-to-point trajectory within a certain period
of time (one second here). One of the motion proles of the end-eector can be given as:
x
y
_ _
=
a
0
a
1
a
2
a
3
0
0 b
0
b
1
b
2
b
3
_ _
1
t
t
2
t
3
t
4
_
_
_
_
_
_
_
_
_
_
_
_
; (30)
where a
i
(i = 0; 1; 2; 3) and b
i
(i = 0; 1; 2; 3) are coecients which can be determined by the
specied conditions at the beginning and the end of the trajectory. The velocities of the end-ef-
fector at the beginning and end of the trajectory are supposed to be zero. If the start-point is
(x
i
; y
i
) and the end-point is (x
f
; y
f
), then the above coecients can be determined as follows:
792 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
a
0
= x
i
;
a
1
= 0;
a
2
= 3(x
i
x
f
);
a
3
= 2(x
i
x
f
);
_

_
(31)
b
0
= y
i
;
b
1
= a;
b
2
= 4(y
i
y
f
) 2a;
b
3
= a 3(y
i
y
f
);
_

_
(32)
where a R is an arbitrary real number and supposed to vary within [)1.0, 1.0] in the following
simulation.
6. Simulation results and discussion
6.1. Parameters
A planar robot with three exible links (as shown in Fig. 2) is used here in the numerical
simulation study. The planar robot manipulator has one redundant degree of freedom.
The cross-section of each link is rectangular and the material used for each link is aluminum. l
e
and l
p
are supposed to be constant, and the end-eector is supposed to be xed at the endpoint.
One element is used for each link in the simulation. The total number of the generalized coor-
dinates nu = 12. Other parameters are given as follows:
length of each link is 250 mm;
height of each link is 5 mm;
width of each link is 8 mm;
elasticity modulus is 7:10 10
10
Pa;
shear modulus is 2:60 10
10
Pa;
Fig. 2. A planar exible redundant manipulator.
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 793
density is 2:71 10
3
kg m
3
;
lumped mass at distal end of each link is 40 g;
lumped mass at endpoint is 20 g;
moment of inertia of base joint is 1:5 10
5
kg m
2
;
moment of inertia of second joint is 1:0 10
5
kg m
2
;
moment of inertia of third joint is 0:5 10
5
kg m
2
;
transmission ratio of each motor is 200:1.
Fig. 3. Comparison of the optimal joint positions.
794 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
For the same q
10
of the planar FRM, there are only two dierent postures for the same end-point
(see Fig. 2). The innite possible postures can be described by q
10
in this case if one of the two
dierent postures is selected throughout in the following numerical simulation. The initial q
10
is
varied within [)1.0, 1.9].
Fig. 4. Comparison of the optimal joint velocities.
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 795
6.2. Simulation results
The simulation of the planar FRM is conducted in the vertical plane with gravity eect. The
gravitational constant is supposed to be 0.98 m s
2
. The trajectory is digitized to one hundred
points, i.e. w = 100. The joint nominal motion planning is conducted according to Eq. (21). The
joint-angle constraint [q
l
; q
u
[ is [5p=6; 5p=6] for each joint, k of each motor is 0.40, 0.20 and 0.10
Fig. 5. Comparison of the optimal joint accelerations.
796 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
respectively, the nominal allowed torque of each motor is 1.0, 0.8 and 0.6 Nm, respectively, and
d
a
is supposed to be (0:001m 0:001 m)
T
.
The optimization software OPT-1 [22] is used here. The initial parameters of the optimization
are set to be: initial conguration q
10
= 1:0 rad, the initial payload mass is m
p
= 0:02 kg, and the
moment of inertia of the payload corresponding to m
p
= 0:02 kg is I
p
= 0:10 10
5
kg m
2
.
The process of searching for the optimum trajectory converges after ve iterations. The sim-
ulation results are graphically illustrated.
The optimal joint angle, joint velocity and joint acceleration are compared in Figs. 35. It is
noticed that the obtained conguration (i.e., joint angles together with their derivatives) of the
optimal result is quite dierent from that of the initial and intermediate in Fig. 3. This means that
the redundancy of the robot is important in increasing the payload capacity in this example.
The endpoint vibration deformation of the FRM is shown in Fig. 6. The deformation changes
rapidly within the upper and lower bounds of constraint (0:001 m). The computed torques are
plotted in Fig. 7. It can be found that the torques associated with the exible vibration uctuate
seriously within the bounds. The optimal, initial and intermediate trajectories are compared in
Fig. 8. The optimal solution at each iteration is summarized in Fig. 9. For the given example, a
payload of m
p
= 0:253 kg with I
p
= 1:265 10
5
kg m
2
is found to be the maximum dynamic
payload that the FRM can carry over the point-to-point trajectory while not violating any of the
constraints.
7. Conclusions
High payload-to-mass ratio is one of the advantages of exible robot manipulators. For a
given point-to-point task, the maximum dynamic payload-carrying capacity of a FRM is
Fig. 6. Endpoint deformation with optimal trajectory.
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 797
established in the paper. A nite-element model is employed for describing the dynamics of the
system considering the added payload and the exibility of the FRM. Dierent from rigid-body
manipulators, vibration deformation, which is due to the lightweight and high operating ac-
celerations, is considered as constraints for the system optimization. The kinematic redundancy
is integrated into the optimization. The maximum-dynamic-payload trajectory is determined
through optimization, with a planar FRM used as an example. This trajectory synthesis method
Fig. 7. The optimal torque against torque for maximum dynamic payload.
798 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800
is useful especially for the FRM performing repetitively point-to-point payload-carrying oper-
ations.
Potential application of the presented method to general robot arms may also be possible to
improve the resolution of these arms, which suer from joint's and/or link's elasticity.
Acknowledgements
The work is supported by the Hong Kong Research Grant Council (project #9040306).
Fig. 8. Optimal trajectory for maximum payload.
Fig. 9. Optimal payload computed at each iteration.
S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800 799
References
[1] K. Desoyer, P. Kopacek, P. Lugner, I. Troch, Flexible robot a survey, in: Proceedings of the IFAC, Vienna,
Austria, 1986, pp. 2334.
[2] P.E. Gaultier, W.L.Cleghorn, Modeling of exible manipulator dynamics: a literature survey, in: In Proceedings of
the First Nat. Appl. Mechanism and Robot Conference, Cincinnati, OH, 1989, pp. 2c3.110.
[3] W.J. Book, Modeling, design, and control of exible manipulator arms: a tutorial review, in: Proceedings of the
29th IEEE Conference on Decision and Control, 1990, pp. 500506.
[4] W. Yim, Jichun Zuang, S.N. Singh, Experimental two-axis vibration suppression and control of a exible robot
arm, J. Robotic Syst. 10 (3) (1993) 321343.
[5] N.C. Singer, W.P. Seering, Using a causal shaping techniques to reduce robot vibration, IEEE Conf. Robotics
Autom. (1988) 14341439.
[6] W.E. Singhose, W.P. Seering, N.C. Singer, Shaping inputs to reduce vibration: a vector diagram approach, IEEE
Proc. Conf. Robotics Autom. (1990) 922927.
[7] K.L. Hillsley, S. Yurkovich, Vibration control of a two link exible robot, IEEE Proc. Conf. Robotics Autom.
(1991) 21212126.
[8] A. Konno, M. Uchiyama, Y. Kito, M. Murakami, Conguration-dependent vibration controllability of exible-
link manipulators, Int. J. Robotics Res. 16 (4) (1997) 567576.
[9] L.T. Wang, B. Ravani, Dynamic load carying capacity of mechanical manipulators part I: problem formulation,
Trans. ASME J. Dynamic Syst., Meas., Control 110 (1988) 4652.
[10] L.T. Wang, B. Ravani, Dynamic load carrying capacity of mechanical manipulators part II: computational
procedure and applications, Trans. ASME J. Dynamic Syst., Meas., Control 110 (1988) 5361.
[11] T.R. Parks, H.A. Pak, Eect of payload on the dynamics of a exible manipulator-modeling for control, ASME J.
Dynamic Syst., Meas., Control 113 (1991) 409418.
[12] Q. Tu, J. Rastegar, Eects of payload on the vibrational excitation of robot manipulators during motion, Proc.
ASME Conf. Mach. Elements Mach. Dyn. (1994) 1724.
[13] M.H. Korayem, A. Basu, Formulation and numerical solution of elastic robot dynamic motion with maximum
load carrying capacities, Robotica 12 (1994) 253261.
[14] D.N. Nenchev, Redundancy resolution through local optimization: a review, J. Robotic Syst. 6 (6) (1989) 769798.
[15] B. Siciliano, Kinematic control of redundant robot manipulators: a tutorial, J. Intelligent Robotic Syst. 3 (1990)
201212.
[16] L.A. Nguyen, I.D. Walker, R.J.P. Degueiredo, Dynamic control of exible kinematically redundant robot
manipulators, IEEE Trans. Robotics Autom. 8 (6) (1992) 759767.
[17] S.G. Yue, Redundant robot manipulators with joints and link exibility I: dynamic motion planning for minimum
end eector deformation, Mech. Mach. Theory 33 (12) (1998) 103113.
[18] S.G. Yue, Weak-vibration congurations for exible robot manipulators with kinematic redundancy, Mech. Mach.
Theory 35 (2000) 165178.
[19] S. Kim, Y. Park, Self-motion utilization for reducing vibration of a structurally exible redundant robot
manipulator system, Robotica 10 (1998) 669677.
[20] S.G. Yue, Y.Q. Yu, S.X. Bai, Flexible rotor beam element for robot manipulators with link and joint exibility,
Mech. Mach. Theory 32 (2) (1997) 209219.
[21] W.L. Cleghorn, R.G. Fenton, B. Tabarrk, Finite element analysis of high-speed exible mechanisms, Mech. Mach.
Theory 16 (4) (1981) 407424.
[22] Z. Yu, J. Zhou, Theory and introduction for optimization software OPB-1, Mechanical Engineering Publishing
House, 1989, Beijing (in Chinese).
800 S. Yue et al. / Mechanism and Machine Theory 36 (2001) 785800

S-ar putea să vă placă și