You are on page 1of 6

Proceedings of the 40th IEEE

Conference on Decision and Control


Orlando, Florida USA, December 2001

FrA01-4

N o n l i n e a r C o u p l i n g C o n t r o l Laws for a 3 - D O F O v e r h e a d
Crane System 1
Y. Fang zx, W. E. Dixon t, D. M. Dawson A, and E. Zergeroglu*
ADepartment

of Electrical & Computer Engineering, Clemson University, Clemson, SC 29634-0915, [yfang, ddawson]@clemson.edu

~Oak Ridge National Laboratory, P.O. Box 2008, Bldg. 7601, Mailstop 6305, Oak Ridge, TN 37831, dixonwe@ORNL.GOV
SLucent Technologies, Bell Lab Innovations Optical Fiber Solutions, 50 Halls Rd, Sturbridge MA, 01566, ezerger@lucent.com

Abstract: In this paper, we consider the regulation control


problem for a three-degree-of-freedom (3-DOF), underactuated
overhead crane system. Motivated by recent passivity-based controllers for underactuated systems, we design several controllers
that asymptotically 'regulate the planar gantry position and the
payload angle. Specifically, utilizing LaSalle's Invariant Set Theorem, we first illustrate how a simple proportional-derivative
(PD) controller can be utilized to asymptotically regulate the
overhead crane system. Motivated by the desire to achieve i'mproved transient performance, we then design two nonlinear controllers that increase the coupling between the planar gantry position and the payload angle.
1 Introduction
Precise payload positioning by an overhead crane (especially when performed by a operator using only visual feedback to position the payload) is difficult due to the fact that
the payload can exhibit a pendulum-like swinging motion.
These payload swings can result in several performance and
safety concerns including: i) damage to the payload (e.g.,
spillage or breakage), ii) damage to the surrounding environment or personnel, and iii) large internal forces that can
result in reduced payload carrying capacity or premature
failure of stressed parts. Motivated by the desire to achieve
fast and precise payload positioning while mitigating the
above performance and safety concerns, several researchers
have developed various controllers for overhead crane systems. For example, Yu et al. [19] utilized a time-scale separation approach to control a two degree-of-freedom (2-DOF)
overhead crane system; however, an approximate linearized
model of the crane was utilized to facilitate the construction of the error systems. In [17], Yashida et al. proposed
a saturating control law based on a guaranteed cost control
method for a linearized version of the 2-DOF crane system
dynamics. Martindale et al. [10] utilized an approximate
crane model to develop exact model knowledge and adaptive controllers while Butler et al. [2] exploited a modal
decomposition technique to develop an adaptive controller.
In [3], Chung and Hauser designed a nonlinear controller
for regulating the swinging energy of the payload.
Several researchers have also examined the control problem
for 3-DOF overhead crane systems. Specifically, Moustafa
1This research was performed in part by a Eugene P. Wigner
Fellow and staff member at the Oak Ridge National Laboratory,
managed by UT-Battelle, LLC, for the U.S. Department of Energy under contract DE-AC05-00OR22725 and is supported in
part by the U.S. NSF Grants DMI-9457967, DMI-9813213, EPS9630167, ONR Grant N00014-99-1-0589, a DOC Grant, and an
ARO Automotive Center Grant.

0-7803-7061-9/01/$10.00 2001 IEEE

3766

and Ebeid [11] derived the nonlinear dynamic model for a


3-DOF overhead crane and then utilized a standard linear
feedback controller based on a linearized state space model.
In [12], Noakes and Jansen developed a generalized input
shaping approach for the linearized crane dynamics that
exploited a notch filtering technique to control the motion
of the bridge/trolley of a 3-DOF overhead crane system.
More recently, Lee [8] developed a nonlinear model for 3DOF overhead cranes based on a new 2-DOF swing angle
definition. Based on this nonlinear model, Lee then developed an anti-swing control law for the decoupled linearized
dynamics. In [13], Sakawa and Sano derived a nonlinear
model for a 3-DOF crane system that was subsequently
linearized to facilitate the development of a control scheme
that first transferred the load to a position near the equilibrium point using an open-loop controller and then utilized
a linear feedback controller to stabilize the payload about
the equilibrium point.
One of the limiting factors associated with the above overhead crane control designs is that the system nonlinearities
are often excluded from the closed-loop error system design
and stability analysis. To overcome this drawback, several
researchers have investigated control approaches that account for the nonlinear dynamics of overhead cranes and
similar systems. For example in [15], Teel utilized saturation functions to develop an output feedback controller
which achieves a robust, semi-global stability result for the
ball-and-beam control problem. In [1], Burg et aI. transformed the nonlinear crane dynamics into a structure that
resembled the ball-and-beam problem and then adopted the
research efforts of [15] to achieve asymptotic positioning
from a large set of initial conditions. More recently, Fantoni
et al. [6] and Lozano et al. [9] proposed passivity-based controllers for the inverted pendulum and the pendubot (i.e.,
an inverted pendulum-like robot with an unactuated second
link) based on the paradigm of driving the underactuated
system to a homoclinic orbit using an energy-based nonlinear controller and then switching to a linear controller to
stabilize the system around its unstable equilibrium point.
Using similar stability analysis techniques, Collado et al.
[4] proposed a proportional-derivative (PD) controller for
the overhead crane problem. In [7], Kiss et al. developed a
PD controller for a vertical crane-winch system that only requires the measurement of the winch angle and its derivative
rather than a cable angle measurement. In [18], Yoshida
developed a nonlinear energy-based controller to damp out
the pendulum oscillations of a 2-DOF overhead crane system despite amplitude constraints on the trolly position.

Recently, Fang et al. [5] developed several energy based


controllers for a 2-DOF overhead crane system in which
additional nonlinear terms were injected into the controller
to increase the coupling between the gantry position and
the payload position to provide for improved transient response.

Vrnla = - m p L sin 0 sin 0 + m p L cos 0 cos


gin14 -- m p L c o s O cos0 - m p L sin 0 sin

Vm2a = - m p L sin 0 cos 0 - m p L cos 0 sin


Vm24 -- - r n p L cos 0 sin 0 - m p L sin 0 cos
Vm34 = - m p L 2 sin 0 cos t)

In this paper, we utilize a similar approach as provided by


[5, 9] to develop several controllers for the 3-DOF overhead
crane system. Specifically, utilizing LaSalle's Invariant Set
Theorem, we illustrate how a simple PD controller can be
utilized to asymptotically regulate the overhead crane dynamics. Motivated by the desire for improved transient
response, we design two nonlinear energy-based coupling
control laws that increase the coupling between the pendulum position and the gantry position. Simulation results
are provided, which illustrate that the increased coupling
of the nonlinear controllers results in improved transient
response (e.g., reduced overshoot and faster settling time)
over the P D control law. The paper is organized as follows.
In Section 2, we present the nonlinear dynamic model of
the overhead crane system and in Section 3, we rewrite
the open-loop system into a more convenient form. In Section 4, we develop a P D controller and two nonlinear controllers and examine the stability of the controllers through
a Lyapunov-like stability analysis. The performance of the
proposed controllers is illustrated through simulation results presented in Section 5. Concluding remarks are given
in Section 6.

Vm43 "- m p L 2 sin 0 cos 0


Vm44 = mpL2 sin 0 cos 0i)
G=[0

=[F~

The dynamic model for a three degree-of-freedom (3-DOF)


overhead crane system (see Figure 1) is assumed to have
the following form [11]

M(q)~ + Vm(q, gl)q + G(q) = u

y 0 1~

0
m31
m41

m22
m32
m42

m23
m33
0

roll = m p + mr + me,

m24
0
m44

~T (11]/I(q) - Vm(q,~l)) ~ = O

m24 = - m p L sin 0 sin

rn31 = m p L cos 0 sin ,

m32 = m p L cos 0 cos

where/~/(q) represents the time derivative of M(q) and that


the inertia matrix M(q) can be upper and lower bounded
by the following inequalities
~___k2 II~[I2

v~ ~ R 4

(8)

where kl, k~ E R 1 are positive bounding constants. In a


similar manner as in [1] and [10], we assume that the dynamic model given in (1) has the following characteristics.

(2)

(3)

A s s u m p t i o n 3: The gantry mass and the length of the


connecting rod are exactly known.
A s s u m p t i o n 4: The ball joint that connects the payload
link to the gantry is frictionless and that this joint
does not rotate about the connecting rod (i.e., the
payload does not rotate about the rod axis).
A s s u m p t i o n 5: The angular position of the payload mass
is restricted according to following inequality

-~r < O(t) < 7r

(9)

where O(t) is measured from the vertical position (see


Figure 1).

m 4 1 = m p L sin Ocos

m42 = - m p L sin 0 sin ,

I O
0
0
0

(7)

A s s u m p t i o n 2: The angular position and velocity of the


payload and the planar position and velocity of the
gantry are measurable.

m22=mp+rnc

m23 = m p L cos 0 cos ,

Vm =

V~ E R4

rnl3 = m p L cos 0 sin

ml4 = m p L sin Ocos ,

m33=mpL 2+I,

(6)

A s s u m p t i o n 1: The payload and the gantry are connected by a massless, rigid link.

where x(t) E R 1 denotes the gantry position along the XCoordinate axis, y(t) E R 1 denotes the gantry position
along the Y-Coordinate axis, O(t) E R 1 denotes the payload angle with respect to the vertical, (t) E R 1 denotes
the projection of the payload angle along the X-Coordinate
axis, and M(q) E R 4x4, Y,~(q,(t) E R 4x4, G(q) E R 4, and
u(t) E R 4 are defined as follows
M --

F~ 0 0] ~

(5)

(1)

where q(t) E R 4 is defined as follows

q=[~

0] T

where m p , m r , m c E R 1 represent the payload mass, rail


mass, and cart mass, respectively, I E R 1 denotes the moment of inertia of the payload, L E R 1 represents the length
of the crane rod, g E R 1 represents the gravity effects, and
F~(t),Fy(t) E R 1 represent the control force inputs acting
on the cart and rail, respectively. Based on the structure of
M(q) and V,~(q, q) given in (3) and (4), it is straightforward
to show that the following skew-symmetric relationship is
satisfied

]gl II~ll2 ~___~TM(q)~


2 Dynamic Model

mpgLsinO

m44 = m p L 2 sin 2 0 + I

0 Vm13
0 Vm23
0
0
0 Vm43

Vml4

Vrn24
gm34
Vm44

(4)

3767

R e m a r k 1 Note that the model given by (1) could be modified to include other dynamic effects associated with the
gantry dynamics (e.g., gantry friction, viscous damping coej~cients, mass m o m e n t of inertia o.f the gantry and rail
motors, etc.); however, these additional dynamic effects
were not included in the model since these effects can be
directly cancelled by the controller.

w2 =

m c

~ ,

....

mpL [ mp + mr + me) I + (mr + me)mpL 2sin 2 0]


[d2 (mpL2.sin 2 0 + I) + p2 (mpL 2 + / ) ] s i n 0 cos
+2rnpLIO cos 0 sin (mp + rnr + mc) (mpL 2 + I)
-2mpLIO cos 0 sin em2pL2 cos 2 0
+mpgLsinO [(mr + m ~ ) mvL
2 3 sin 2 0 cos 0 cos
+(mr + m e +rnp)mpLIcosOcos].
(18)

In order to write the open-loop dynamics given in (12) and


(13) in a more compact form for the subsequent control
development and stability analysis, we define the auxiliary
signal r(t) E R 2 as follows

/oi

r = [x y]T.

I / %%
,,- -@ ,.

X Y Z : Fixed
coordinate s y s t e m

After taking the second time-derivative of r(t) and then


utilizing the expressions given in (12)-(18), we can rewrite
the open-loop dynamics given in (12) and (13) as follows

mp

=
Figure

(19)

Ix~ ]

= det(M)

(PF+W)

(20)

where P(q) C R 2x and W(q, (1) E R 2 are defined as follows

1: 3 - D O F O v e r h e a d C r a n e S y s t e m

3 Open-Loop System Development


and F(t) E R 2 is defined below
To express (1) in a form that facilitates the subsequent
control development and stability analysis, we premultiply
both sides of (1) by M-~(q) to obtain the following expression
q = M -~ ( u - V m ~ - G)
(10)
where M-l(q) E R 4x4 is guaranteed to exist due to the
fact that the determinant of M(q), denoted by det(M), is
a positive function as shown below
det(M) = 12 (mp + me) (mp + mr + rn~) +

+ m )mo (1 + sin 0) +

mpL2I (mp + mc) mp sin 2 0+


mpL2Imrmp (sin 2 + 2 sin 2 0 cos 2 ) +
2 4
mpL
sin 20rn~mp sin 2 0 cos 2 4)+
2 4
mpL
sin 20m~ (mr + me +mp sin 2 0). (11)
After substituting (3), (4), (5), (6) and the expression for
M - l ( q ) into (10) and performing some algebraic manipulation, we can rewrite (10) as follows
1
= d e t t M ) (PllF.; + p12Fu + wz)
(12)

# =,

. . . . . (p~2F~ + p22Fy + w2)


(13)
ae~,lvl )
where the measurable auxiliary terms pll (q), p12(q), p22(q),
wl(q, (:1), w2(q, 4) 6 n I a r e defined as f o l l o w s
2 2 I (sin 2 + 2sin 2 0cos 2 ) + mpI 2 +
Pll -- mpL
mempL 4 sin 2 0 + mcmpL2I (1 + sin 2 0)
+mcI 2 + m3L 4 cos 2 esin 4 0
P12 = - m p3L 4 sin cos sin4 0 2 2I sin cos (sin2 0 - cos 2 O)
mpL
3 4
P22= mpL sin a0sin 2 + ( m v + m r + m e ) I 2 +
m2L2I [1 + (sin2 0 - cos 2 0)sin 2 ] +
(mr + m~) m2pL4 sin 2 0+

(mr + m~)mpL2I (1 + sin 2 O)

wl =

F = [F~

F~] ~ .

(22)

Given the expressions in (14)-(16), it is straightforward to


prove that
Pll

> 0

and pllp22-p212 > mp (rap -F 'mr -4-mc)14. (23)

From the expressions given in (21) and (23), we can see that
P(q) is positive-definite, symmetric, and invertible, where
the inverse of P(q), denoted by p-l(q), is also positivedefinite and symmetric.
'
To facilitate the subsequent Lyapunov-based control design,
we will utilize the energy of the overhead system, denoted
by E(q, (1) E R 1, and defined as follows
lq T

E(q, (1) = -~

M(q)(l + mpgn(1 - cos(0)) > 0.

(24)

After taking the time derivative of (24), substituting (1)


for M(q)~l(t), and canceling common terms, we obtain the
following expression for the time derivative of E(q, (t)
E=~F

(25)

where (3), (4), and (7) were utilized.


4 Control Design and Analysis

(14)

(15)

Our control objective is the regulation of the planar gantry


position of the overhead crane to a constant desired position, denoted by rd E R 2, which is explicitly defined as
rd=[Xd

(16)

mpL sin 0 sin 4) [(mp + me) I + mpmcL 2 sin 2 0]

0+ , ) + 0

W]~.

In addition, the payload angle O(t) must also be regulated


to zero. To quantify the control objective of regulating the
overhead crane to a constant desired position, we define a
gantry position error signal e(t) E R 2 as follows
~(t) = ~ - ~ .

-2ImpLO~bcosO cos (rap + rnc) I+


m2 gL 2 sin 0 cos 0 sin [(rap + m~) I + mvm~L 2 sin 2 0]

+2/m, n0~o~0cosem, L~ (,~o +,n~si, ~0) (17)


3768

(26)

(27)

In the subsequent control development, we will design a


proportional-derivative control law and two nonlinear controllers to achieve the above control objective.

R e m a r k 2 The control objective is defined in terms of regulating the gantry position and the angle of the payload with
the vertical. The problem of regulating the projection of the
payload angle along the X-Coordinate axis, denoted by (t),
is not required. That is, if the payload angle, denoted by
O(t), is regulated to zero, then we can see from Figure 1
that the payload has been regulated to the desired location.
4.1 P r o p o r t i o n a l - D e r i v a t i v e
Control Law
Based on the subsequent stability analysis, we design the
following proportional-derivative (PD) control law

F -- - k p e - kdi~
k~

(2s)

where kd, kF., kp E R 1 are positive constant control gains.


T h e o r e m 1 The controller given in (28) ensures asymptotic regulation of the overhead crane system in the sense
that
lim ( x ( t ) y ( t ) O ( t ) ) = (

Xd

Yd

0 )

Based on (36), it is clear t h a t E(q, (t) and e(t) are constant,


and hence, from (28) and (34), it is clear that F(t) is constant. To complete the proof, we must analyze the stability
of the system for the case when 0 = 0 and when ~ ~ 0.
In this analysis, given in Appendix A, we prove the result
given in (29) under the proposition that 0 = 0 and that the
proposition that 0 # 0 leads to contradictions, and hence,
is an invalid proposition.
R e m a r k 3 In the previous stability analysis, we have
shown that the control objective is met and that all signals
in the dynamics and the controller re'lnain bounded for all
time except for the signal (t) (Note that by assumption, the
payload angle, denoted by O(t), is assumed to be bounded).
We note that the boundedness of (t) is insignificant from
a theoretical point of view since (t) only appears in the dynamics and control as arguments of trigometric functions.
We also note that the simulation results for all of the controllers indicate that the signal (t) remains well-behaved
and is driven to constant value.

(29)

t----~oO

where Xd and Yd were defined in (26).


P r o o f : To prove (29), we define a nonnegative function
V1 (t) E R 1 as follows
1 k p e T e.
Vl = k s E + -~

(30)

After taking the time derivative of (30) and then substituting (25) and the time derivative of (27) into the resulting
expression, we can rewrite (30) as follows

?~ = T (kEF + kpe).

(31)

After substituting ( 2 8 ) i n t o (31) for F(t) and then cancelling common terms, we obtain the following expression

y~ =

-kS.

(32)

Based on the expressions given in (8), (24), (27), (30)


and (32), it is clear that the origin of the closed-loop
system is stable in the sense of Lyapunov [14] and that
r(t), e(t), (t(t) E .~. Based on the fact that r(t), e(t), (t(t)
E E ~ , we can utilize (2) and (19) to prove that x(t), it(t),
y(t), y(t), (t), b(t), (t) E ~ . Given that e(t), (t) E ~,
it is clear from (28) that F(t) E 12~. Finally, from (6) and
(22), we c a n prove that F,(t), Fu(t), u(t) E 12~.
Based on the fact that all of the closed-loop signals remain
bounded, we can now employ LaSalle's Invariance Theorem
to prove (29). To this end, we define F as the set of all points
where
i~'~ = 0 .
(33)
In the set F, it is clear from (32) and (33) that

~(t) = 0

~(t) = 0,

(34)

and hence, we can conclude from (19), (30), (33), and (34)
that x(t), y(t), and ]/1 (t) are constant, and that
2(t) = 0

9(t) = 0.

(35)

Furthermore, from (25), (27), and (34), it is clear that


/~(q, (~) -- ~(t) - 0.

' (36)

3769

R e m a r k 4 Heuristically, the only way for the energy from


the payload motion to be dissipated is through the coupling between the payload dynamics and the gantry dynamics. That is, a PD feedback loop at the gantry creates an
artificial spring/damper system which absorbs the payload
energy through the natural gantry/payload coupling, however, from our experience with many control experiments
on overhead crane testbeds, we believe that a PD feedback
loop at the gantry will always provide poor performance because if the gantry friction is not compensated for perfectly
(and it never will be), then the uncompensated gantry friction effects tend to retard the natural co~,pling between the
gantry/payload dynamics, and hence, p'l'evcnt payload energy from being dissipated by the PD .feedback loop at the
gantry. In the following sections, we develop controllers
that may improve the performance o.f the PD feedback loop
at the gantry, due to the incorporatiol~ o.f additional nonlinear terms in the control law that depend on the payload dynamics. Although the subsequent contr'ollers yield the same
stability result as the previous controller, 'we believe that the
increase in gantry/payload coupling due to the additional
nonlinear terms will result in improved performance when
compared to the simple gantry PD controller.
4.2 E 2 C o u p l i n g C o n t r o l L a w
Based on previous work presented in [9] for ~ 2-DOF inverted pendulum, we design the following E 2 coupling control law 1
F = [gt]

- kdr - kpe - d ~ / ) ~

IV

(37)

where gt(t) E R 2x2 is an auxiliary positive-definite, invertible matrix 2 defined as follows


k.
gt = kEEI2 + det(J~,~ 1)
(38)
1The control strategy is called an E 2 coupling control law
because its structure is motivated by a SClu~r~l energy term in
the Lyapunov function and an additional ga.~t.ry squared velocity
term in the Lyapunov function.
2Since P and I2 are positive definite m~.triccs, and kE, kv,
E(q, q), and det(M(q)) are positive scalals, it is clear that fl(t)
is positive definite and invertible.

kE, kp, kd, kv E R 1 are positive c o n s t a n t control gains, /2


denotes the standard 2 x 2 identity matrix, and det(M),
P(q), and W(q, (l) were defined in (11) and (21).

T h e o r e m 2 The controller given in (37) ensures asymptotic regulation of the overhead crane system in the sense
that
lim ( x ( t )

~----*00

y(t)

8(t) ) =

( Xd

Yd

T a b l e 1" Control Gains

PD
Controller

E 2 Coupling
Control Law

102
45
1
No

125.3
50
0.001
50

kd

0 )

ks
kv

Gantry Energy
Coupling
Control Law
350
120
0.4
0.6

where Xd and Yd were defined in (26).

Proof: See Appendix B.


Cart Position y(t)

Cart Position x(t)

4.3 Gantry Kinetic Energy Coupling Control Law


To illustrate how additional controllers can also be derived,
we design the following nonlinear coupling control law 3
--kdi" - kpe - kvP - 1 W - ~ k1 v ( d ( d e t ( M ) p - Z ) )
F =

(39)

k s + kv

1.5

1.5

Z1
0.5

05

10

20
Angle e(t)

30

10

20
Angle (t)

30

~o

30
'

2'0

3?

40

10

where kE, kp, kd, and kv E R 1 are positive constant control


gains, and Get(M), P(q), and W(q,(t) were defined in (11)
and (21).

x_i!
10

T h e o r e m 3 The controller given in (39) ensures asymptotic regulation of the overhead crane system in the sense
that

40

10

600

2o
30
Force Input Fx(t)

40

'"

lo

Force Input Fy(t)

600

2OO

lim ( x ( t )

O(t) ) =

y(t)

( Xd

t--*O0

0 )

(40)

-2000

,o

io

3'o

01

40

Time[sec]

where

Xd

,0

Time[sec]

and Yd were defined in (26).

F i g u r e 2: Results for the P D Controller

Proof: See Appendix C.


5 Simulation Results

Cart Position x(t)

To illustrate the performance of the proposed controllers,

Cart Position y(t)


4

15

we simulated the 3-DOF


crane system given in (I), where
the crane parameters were selected as follows

mp = 160 [kg],
I = 1.5 [kg.m2],

m~ = 23 [kg],
L = 2.5 [m]

m~ = 190 [kg],
(41)

and the desired position of the crane gantry was selected as


follows
[ Xd Yd ] T = [ 10 3 ]T.
(42)
For each of the simulations, the initial conditions were set
to zero and the control gains were tuned until the best performance was achieved. The resulting control gains from
each controller is given in Table 1.

,,

0 v ....
0

10

, o.:L

20
Angle O(t)

~0~

30

0.2

10

20
Angle (t)

30

40

Force 120putFy(t) 30

40

2
1.5

.....
-0.11!,
-02
0

0.5
'

20
30
10 FForce Input Fx(t)

40

10

'

200 /

E 1[1~

The resulting gantry position error, payload angle, and the


input force are shown in Figure 2 for the PD control law,
Figure 3 for the E 2 coupling control law, and Figure 4 for
the gantry kinetic energy coupling control law.
3The control strategy is called a gantry kinetic energy coupling control law because its structure is derived from an additional gantry kinetic energy-like term in the Lyapunov function.

3770

lO

2'0

Tirne[sec]

io

4o

~o

~'o

3'o

Time[sec]

F i g u r e 3: Results for the E 2 Coupling Control Law

Cart Positionx(t)

Cart Posilion y(l)

15

4
3

10

5
0
0
0.4

10

20
Angleo(t)

30

10 Force 120ut_Fx(t) 30
'

10

20
Angle(t)

30

021.il

20001

"-'0 ~

0'5I
0
0

lO Fo,~e12nOutFy(t)

500 ~ I .

"

/
~, 1 k
-500
0

10

20
Time[sec]

30

40

-1000

10

20
Time[sec]

30

F i g u r e 4" Results for the Gantry Kinetic Energy Coupling


Control Law

6 Conclusion
In this paper, we presented three controllers for an overhead crane system. By utilizing a Lyapunov-based stability analysis along with LaSalle's Invariance Theorem, we
proved asymptotic regulation of the gan.try and payload
position for a PD controller and two nonlinear controllers.
Simulation results were utilized to demonstrate that the
increased coupling between the gantry and payload that results from the additional nonlinear feedback terms in the
nonlinear coupling control laws, resulted in improved transient response. Future work will focus on comparing the
performance of the P D controller with the nonlinear coupling control laws through experimental results obtained
from an overhead crane system with a gantry that moves in
a 2-DOF Cartesian plane.
References

[1]
T. Burg, D. Dawson, C. Rahn and W. Rhodes, "Nonlinear Control of an Overhead Crane via the Saturating
Control Approach of Teel", Proc. IEEE Int. Conf. Robotics
and Automation, pp. 3155-3160, 1996.
[2]
H. Butler, G. Honderd, and J. Van Amerongen,
"Model Reference Adaptive Control of a Gantry Crane
Scale Model", IEEE Control Systems Magazine, pp. 57-62,
January 1991.
[3]
C. Chung and J. Hauser, "Nonlinear Control of a
Swinging Pendulum", Automatica, Vol. 31, No. 6, pp. 851862, 1995.
[4]
J. Collado, R. Lozano and I. Fantoni, "Control of
Convey-crane Based on Passivity", Proc. American Control
Conference, pp.1260-1264, 2000.
[5]
Y. Fang, E. Zergeroglu, W. E. Dixon, and D. M. Dawson, "Nonlinear Coupling Control Laws for an Overhead
Crane System", Proc. of the IEEE Conference on Control
Applications, in press.

3771

[6]
I. Fantoni, R. Lozano, and M. W. Spong, "Energy
Based Control of the Pendubot", IEEE Transactions on
Automatic Control, Vol. 45, No. 4, pp. 725-729, 2000.
[7]
B. Kiss, J. Levine, and P. Mullhaupt, "A Simple Output Feedback PD Controller for Nonlinear Cranes", Proc.
of the Conference on Decision and Control, pp. 5097-5101,
2000.
[8]
H. Lee, "Modeling and Control of a ThreeDimensional Overhead Cranes", ASME Trans. on Dynamic
Systems, Measurement, and Control, Vol. 120, pp. 471-476,
1998.
[9]
R. Lozano, I. Fantoni, D. J. Block, "Stabilization of
the Inverted Pendulum Around Its Homoclinic Orbit", Systerns & Control Letters, Vol. 40, No. 3, pp. 197-204, 2000.
[10] S. C. Martindale, D. M. Dawson, J. Zhu, and C.
Rahn, "Approximate Nonlinear Control for a Two degree of
Freedom Overhead Crane: Theory and Experimentation",
Proc. American Control Conference, pp. 301-305, 1995.
[11] K . A . F . Moustafa and A. M. Ebeid, "Nonlinear Modeling and Control of Overhead Crane Load Sway", ASME
Trans. on Dynamic Systems, Measurement, and Control,
Vol. 110, pp. 266-271, 1988.
[12] M . W . Noakes and J. F. Jansen, "Ceneralized Inputs for Damped-Vibration Control of S~tspended Payloads", Computers and Electrical Engineering, An International
Journal, 1991.
[13] Y. Sakawa and H. Sano, "Nonlinear Model and Linear Robust Control of Overhead Traveling Cranes", Nonlinear Analysis, Vol. 30, Issue 4, pp.2197-2207, 1997.
[14] J . J . E . Slotine and W. Li, Applied Nonlinear Control,
Englewood Cliff, NJ: Prentice Hall, Inc., 1991.
[15] A.R. Teel, "Semi-global Stabilization of the 'Ball and
Beam' Using 'Output' Feedback", Proc. American Control
Conference, pp. 2577-2581, 1993.
[16] Q. Wei, W. P. Dayawansa and W. S. Levine, "Nonlinear Controller for An Inverted Pendulum Having Restricted
Travel", Automatica, Vol. 31, No. 6, pp. 841-850, 1995.
[17] K. Yoshida and H. Kawabe, "A Design of Saturating
Control with a Guaranteed Cost and Its Application to the
Crane Control System", IEEE Transactions on Automatic
Control, Vol. 37, No. 1, pp. 121-127, 1992.
[18] K. Yoshida, "Nonlinear Controller Design for a Crane
System with State Constraints", Proc. American Control
Conference, pp. 1277-1283, 1998.
[19] J. Yu, F. L. Lewis, and T. Huang, "Nonlinear Feedback Control of a Gantry Crane", Proc. A~nerican Control
Conference, Seattle, Washington, pp.4310-4315, 1995.
Appendices are availabe upon request.

You might also like