Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/267195182

Nonlinear Modeling and Control of Quay-Side Container Cranes

Article

CITATIONS READS
6 178

3 authors, including:

Ziyad N. Masoud Ali H. Nayfeh


German Jordanian University Virginia Polytechnic Institute and State University
51 PUBLICATIONS   1,925 CITATIONS    798 PUBLICATIONS   32,638 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Geometrical determination of an added mass using M/NEMS resonators View project

Nonlinear Behavior of Systems View project

All content following this page was uploaded by Ziyad N. Masoud on 15 January 2015.

The user has requested enhancement of the downloaded file.


Nonlinear Modeling and Control of Quay-Side Container Cranes

Mohammad F. Daqaq, Ziyad N. Masoud, and Ali H. Nayfeh

Virginia Polytechnic Institute and State University


Department of Engineering Science and Mechanics, MC 0219
Blacksburg, Virginia 24061, USA

ABSTRACT
Gantry cranes are traditionally modeled as a simple pendulum. However, a quay-side container crane has a multi-
cable hoisting mechanism which can not be modeled as a simple pendulum. In this paper, a two-dimensional four-
bar-mechanism model of a container crane is developed. The model is then reduced to a double pendulum with a
kinematic constraint. A nonlinear approximation of the oscillation frequency of the simplified model is developed.
The resulting frequency approximation is used to determine the switching times for a bang-off-bang input-shaping
controller. The performance of the controller is numerically simulated on the full model, and is compared to the
performance of similar controllers, which are based on a nonlinear frequency approximation of a simple pendulum
and a linear frequency approximation of a constraint double pendulum. The sensitivity of the controller performance
to the oscillation frequency is investigated. A delayed-position feedback controller is applied at the end of the transfer
maneuver to eliminate residual oscillations without affecting the shaped commands of the input-shaping controller.

1. INTRODUCTION
Increasing demand on faster, bigger, and efficient cranes to guarantee fast turn-over time steered research into
developing faster and efficient crane controllers that meet safety requirements. Inertial forces on the payload due
to crane commanded motion can cause it to experience large sway oscillations. To avoid exciting these oscillations,
cranes are operated at slower speeds such that the oscillations do not cause safety concerns. This increases the
time and cost of loading and unloading operations.

Traditionally, a gantry crane is modeled as a simple pendulum with a rigid massless hoisting cable. The payload
is lumped with the hook and modeled as a point mass. However, in the case of quay-side container cranes, the
model is significantly different, Figure 1. The actual hoisting mechanism consists typically of a set of four cable
arrangement. The cables are hoisted from four different points on a trolley and are attached on the payload side to
four points on a spreader bar used to lift containers. Performance of controllers designed for gantry cranes is far
from satisfactory when applied to quay-side container cranes due to the lack of a good model that best describes
the dynamics of the crane.

Masoud et al. [1] introduced a two-dimensional four-bar-mechanism model of a container crane. They approximated
the model with a constrained double pendulum. The resulting linear frequency of the double pendulum model was
then used to find the delay and gain of a nonlinear delayed-position feedback controller developed earlier by Henry
et al. [13]. The resulting controller was applied to the full crane model. Their simulations demonstrated a very
successful performance. In this paper, we develop a one step input-shaping controller based on the full model
developed by Masoud et al. [1]. We develop a nonlinear frequency approximation of a constrained double pendulum
that approximates the dynamics of the four-bar-mechanism.

Input shaping is a practical open-loop control strategy for gantry cranes. Controllers using various forms of input
shaping are currently used in ports [2]. This technique moves the crane a known distance along a set path.

Alsop et al. [3] were the first to propose a controller based on input shaping. The controller accelerates the trolley
in steps of constant acceleration then kills the acceleration when the payload reaches zero oscillation angle (after
multiples of a full period). The trolley then coasts at constant speed along the path. A replica of the acceleration pro-
Figure 1: Typical quay-side container crane.

cedure is used in the deceleration stage. They used constant-amplitude acceleration steps, and a linear frequency
approximation of a simple pendulum model. Their results showed very little residual oscillations, while transient
oscillation angles were of order of 10◦ during the acceleration/deceleration stages. Alzinger et al. [4] showed that a
two step acceleration profile results in significant reductions in travel time over a one step acceleration profile. Test-
ing on an actual crane has shown that the two-step acceleration profile can deliver both faster travel and minimal
payload oscillations at the target point. However, any deviation from the prescribed parameters, causes significant
payload oscillations.

Jones et al. [5] used a nonlinear approximation of the oscillation period of a simple pendulum to generate an
analytical expression for the duration of the coasting stage as a function of the amplitude and duration of the constant
acceleration steps. The analytical expression was then used to generate a two-step acceleration profile. Their
computer simulations showed a reduction of residual oscillations to about 0.3◦ . Dadone and VanLandingham [6]
used the method of multiple scales to generate an nonlinear approximation of the oscillation period of a simple
pendulum model. They found that the acceleration profile based on the nonlinear frequency approximation could
damp the residual oscillations two orders of magnitude more than the profile based on the linear approximation. The
enhanced performance of the nonlinear strategy was most pronounced for longer coasting distances and higher
accelerations.

Researchers [7, 8] have extensively studied the possibility of using time delay to control mechanical systems. It has
been noticed that systems with time delays exhibit interesting complex responses. Time delay has the capability of
stabilizing or destabilizing dynamic systems. For this reason, they have been used as simple switches to control
the behavior of systems, either by damping out the oscillations or creating chaotic responses which are sometimes
desirable to secure communication signals [9]. Xu and Chang [10] have investigated the effects of time delayed-
position feedback on the van der Pol duffing oscillator. They used functional analysis to transform the delayed
system into a functional derivative equation to investigate the stability and bifurcations of the equilibrium points.
In another investigation, Nayfeh et al. [11] studied the stability and dynamics of machine cutting-tool by using a
single-degree-of-freedom model that contains linear, quadratic, and cubic time delays.

Cheng et al. [12] were among the first to use time delay to control gantry cranes. They proposed a control strategy
which employs time delay control and feedback linearization to move a crane along a pre-defined path and to
eliminate residual oscillations. Their results showed that the strategy is capable of delivering the payload to its goal
with minimal transient oscillations and almost no residual oscillations.

Despite the effectiveness of the input-shaping controllers, residual oscillations at the end of the transfer operation
are very sensitive to the system parameters. Thus any small deviation from these parameters can amplify the
amplitude of the residual oscillations many orders of magnitude. To eliminate such oscillations, we augment the
input-shaping controller with a delayed-position feedback controller applied at the end of the transfer maneuver.
2. MATHEMATICAL MODELS
2.1. Four-Bar Mechanism Model
The two-dimensional model of a quay-side container crane developed by Masoud at al. [1] is shown in Figure 2.
The container is grabbed using a spreader bar. The spreader bar is then hoisted from the trolley by means of four
cables, two of which are shown in the figure. The operator command f (t) drives the trolley. The length L(t) of the
hoisting cables is also controlled by the operator. The hoisting cables are modeled as two distance constraints. The
specific equations of these constraints are
· ¸
(x + R sin(θ) − 12 wcos(θ) + 12 d − f (t))2 + (y − R cos(θ) − 12 w sin(θ))2 − L(t)2
Φ(q, t) = =0 (1)
(x + R sin(θ) + 21 w cos(θ) − 12 d − f (t))2 + (y − R cos(θ) + 12 w sin(θ))2 − L(t)2

where q = [x, y, θ]T is the generalized coordinates vector.

Using Lagrange multipliers, one can write the mixed differential-algebraic set of constrained equations of motion
as [13] · ¸ · ¸ · A¸
M ΦTq q̈ Q
= (2)
Φq 0 λ γ

where M = diag[m, m, J] is the inertia matrix, and QA = [0, −mg, 0]T is the generalized applied force vector, and

∂Φ
Φq =
∂q
∂ ∂Φ
Φqt = ( )
∂t ∂q (3)
∂2Φ
Φtt =
∂t2
γ = −(Φq q̇)q q̇ − 2Φqt q̇ − Φtt

2.2. Constrained Double Pendulum


For the purpose of controller design, we simplify the four-bar-mechanism model as a double pendulum with a
kinematic constraint relating the angles φ and θ as shown in Figure 3. Point O in Figure 2 is the mid point between
points A and D, and P is the midpoint between points B and C. Using the geometric closing equations of the ABPO

Figure 2: A schematic model of a container crane.


and ODCP loops, the following relations are obtained

d sin φ = w sin(θ + φ) (4)


1
l2 = L2 − (d2 + w2 − 2dw cos θ) (5)
4
We use Euler-Lagrange’s equations to find the equation of motion in terms of φ. Assuming constant cable length,
we substitute Equation (4) and Equation (5) in the system lagrangian. We only use a cubic approximation of the
resulting equation of motion.

φ̈ + ω◦2 φ + c1 φ2 φ̈ + c1 φφ̇2 − c2 f¨φ2 − c3 φ3 + c4 f¨ = 0 (6)

where
s
mg(k3 2 + 2k4 2 + k1 2 k3 R)
ω◦ =
k3 (Jk1 2 + m(k3 − k1 R)2 )
m(4k4 2 + (k1 (1 + k1 )2 − 6k2 )k33 R)
c1 =
k3 2 (Jk1 2 + m(k3 − k1 R)2 )
2mk1 ((3 + 2k1 )k3 k5 R − 2k3 2 (k4 − 3k1 k2 R2 ))
+
k3 2 (Jk1 2 + m(k3 − k1 R)2 )
6Jk1 k2 k3 2 (7)
+ 2 2
k3 (Jk1 + m(k3 − k1 R)2 )
m(6k4 + k3 (k3 − k1 3 R + 6k2 R))
c2 =
2k3 (Jk1 2 + m(k3 − k1 R)2 )
gm(12k4 + k4 (k3 − 24k3 k5 + k1 (k1 3 − 24k2 R))
c3 =
6k3 (Jk1 2 + m(k3 − k1 R)2 )
m(k3 − k1 R)
c4 = 2
Jk1 + m(k3 − k1 R)2

Figure 3: A schematic model of a constrained double pendulum model of a container crane.


and
d−w
k1 =
w
d(d2 − w2 )
k2 = 3
r 6w
1
k3 = L2 − (d − w)2 (8)
4
dw
k4 =
8
dw(d2 − 4L2 + dw + w2 )
k5 =
24((d − w)2 − 4L2 )2

3. NONLINEAR ANALYSIS
3.1. Constrained Double Pendulum
To develop an input-shaping controller, we will find an estimate of the frequency of the of the double pendulum. The
linear frequency ω0 is given in Equation (7). To find a nonlinear frequency approximation we use the method of
multiple scales (Nayfeh [13]). First we scale Equation (6) by introducing a bookkeeping parameter ² which is set to
one at the end. For a constant acceleration f¨ = ξ, we scale both ξ and θ at order ², that is

φ = ²φ ξ = ²ξ (9)

Substituting the scaled parameters into Equation (6) yields

φ̈ + ω◦ 2 φ = −c4 ξ − ²2 (c1 φ2 φ̈ + c1 φφ̇2 − c2 ξφ2 − c3 φ3 ) (10)

The time dependence is expanded into multiple time scales as following

t = T0 + ²T1 + ²2 T2 + O(²3 )
d
= D0 + ²D1 + ²2 D2 + O(²3 ) (11)
dt
d2
= D02 + 2²D0 D1 + ²2 D12 + 2²2 D0 D2 + O(²3 )
dt 2

where Dn = ∂Tn . We then seek a solution in the form

φ(T0 , T1 , T2 ) = φ0 (T0 , T1 , T2 ) + ²φ1 (T0 , T1 , T2 ) + ²2 φ2 (T0 , T1 , T2 ) + O(²3 ) (12)

Substituting Equations (11) and (12) into Equation (10)and equating coefficients of like powers of ² we obtain

O(1) : D02 φ0 + ω◦2 φ0 = −c4 ξ


O(²) : D02 φ1 + ω◦2 φ1 = −2D0 D1 φ0
(13)
O(²2 ) : D02 φ2 + ω◦2 φ2 = −2D0 D1 φ1 − 2D0 D2 φ0
− D12 φ0 − c2 ξφ20 − c1 φ0 (D0 φ0 )2 − c1 φ20 D02 φ0 + c3 φ30

The solution of the O(1) equation can be expressed as


c4 ξ
φ0 (T0 , T1 , T2 ) = A(T1 , T2 )eiω◦ T0 − + cc (14)
ω02

Substituting Equation (14) into O(²) of Equation (13) and eliminating the terms that lead to secular terms, we obtain

∂A
=0 ⇒ A = A(T2 ) (15)
∂T1
5
Numerical Solution
Nonlinear Frequency
Linear Frequency

−5

Angle [deg]
−10

−15

−20
0 2 4 6 8 10 12 14 16 18 20
Time [sec]

Figure 4: Time series of the payload sway angle for the constrained double pendulum model of a container crane. The results are obtained for
L = 17.5 m, R = 2.5 m, d = 0.282 m, w = 0.141 m, ξ = 1.75 m/s−2 , θ0 = 0, and θ˙0 = 0.

Now, substituting Equations (14) and (15) into O(²2 ) of Equation (13) and eliminating the terms that lead to secular
terms, we obtain the solvability condition
∂A ξ 2 c4 c3 c4
−2iω◦ + 2 (c1 c4 + 3 2 − 2c2 )A + (2ω◦2 c1 + 3c3 )A2 Ā = 0 (16)
∂T2 ω◦ ω◦
Introducing the polar transformation
1
a(T2 )eiβ(T2 ) A(T2 ) = (17)
2
into Equation (16) and separating real and imaginary parts, we obtain the following two modulation equations:
∂a
=0 (18)
∂T2
· ¸ · ¸
∂β ξ2 3c3 c24 2 a2 3c3
= 2c2 c4 − − c1 c4 − + c1 ω◦ (19)
∂T2 2ω◦3 ω◦2 4 2ω◦
We solve Equations (18) and (19) for the nonlinear frequency approximation of the constrained double pendulum
to obtain
· µ ¶ µ ¶¸
3c3 c1 ξ 2 −3c3 c24
ω = ω◦ 1 − a2 + + + 2c c
2 4 − c c2
1 4 (20)
8ω◦2 4 2ω◦4 2ω◦2
Figure 4 shows that the numerical integration and the multiple scale approximate solution closely match, while the
linear frequency approximate solution quickly drifts away from the numerical solution as a result of the inaccurate
frequency approximation.

3.2. Simple Pendulum


To compare controllers based on the double pendulum model with those based on the simple pendulum model, we
develop a nonlinear frequency approximation of the payload oscillations for the traditional simple pendulum model
of the container crane using the same procedure described in section (3.1). We obtain
µ ¶
a2 ξ2
ω = ω◦ 1 − + (21)
16 4L2 ω◦4
p
where ω◦ = g/L is the linear frequency approximation of the payload oscillations of a simple pendulum.

4. CONTROLLERS DESIGN

4.1. Input Shaping Controller


We will use a bang-off-bang control input (acceleration) as shown in Figure 5. The goal here is to achieve zero
residual oscillations at the end of the maneuver while taking into consideration any delays in the system.
We assume that both the cable length and the system delays are known. We follow the procedure developed by
Dadone et al. [6]. First we apply a positive acceleration for the period of a half swing cycle, then we switch off
the acceleration for a period of time (coast mode ξ = 0). To bring the load into complete stop we apply a negative
acceleration taking into account the known system delays.

Let us first denote the known system delays as τs . The length of τs will determine the least amount of time that the
trolley must spend in the coast mode. By design we need to coast for a time which is an odd multiple of half the
period of the sway oscillation Tc , where Tc is the oscillation period in the coast mode. The minimum coast time ∆tc
will depend on τs . Thus, we can define the coast time as
2n + 1
∆tc = Tc n = 0, 1, 2, 3, ... (22)
2
where n, the number of coasting cycles, is µ ¶
1 2τs
n≥ −1 (23)
2 Tc
Let us denote the total distance traveled by the trolley in a full maneuver as S. By simple manipulation (see also
Figure 5) we can write
1 2n + 1
S = ξTa2 + ξTa Tc (24)
4 4
where Ta is the period of the sway oscillation when a constant acceleration is applied (bang mode). Using the linear,
and nonlinear frequency approximations, Equations (20) and (20), we are able to calculate the periods for the bang
and coast modes.

Now we can summarize the control algorithm as following:

• Using the known values for τs and Tc , determine the number of coast cycles n from Equation (23),.

• Solve Equation (24) for ξ and compare it with ξmax . ξmax is the minimum of two values, the first being the
maximum acceleration that the traverse drivers of the container crane can provide.
• Compute the switching times t1 , t2 , and t3 according to the following equations:

Ta 2n + 1
t1 = , t2 = t1 + Tc , t3 = t2 + t1 (25)
2 2

4.2. Delayed-Position Feedback Controller


The delayed-position feedback controller was developed earlier by Henry et al [9]. The controller creates damping
in the system by making the suspension point track the position of the load, by means of a delayed percentage of

Figure 5: A schematic drawing showing the bang-off-bang acceleration profile.


the payload sway. The delayed-position feedback controller takes the following general form

f = f◦ + k̂ sin φτd (26)

where k̂ is the feedback gain of the controller, τd is the delay time, f◦ is the operator input, and φτd = φ(t − τd ).
To study the linear stability of the controller and make a proper choice of the gain-delay combination, we substitute
Equation (26) into Equation (6) and add a linear damping term to account for the damping in the system. Assuming
that the operator input is slowly varying and keeping only linear terms yield to

φ̈ + ω◦2 φ + 2µφ̇ + c4 k̂(φ̈τd + 2µφ̇τd ) = 0 (27)

where µ is a linear damping coefficient. Now we seek a solution of the equation in the form

φ = ceσt cos(ωt + ψ) (28)

where c, σ, ω, and ψ are real constants. Letting c4 k̂ = K, substituting Equation (28) into Equation (27), and setting
the coefficients of cos(ωt + ψ) and sin(ωt + ψ) equal to zero independently, we get the following two equations

K(σ 2 + 2µσ − ω 2 ) sin(ωτd ) − 2Kω(µ + σ) cos(ωτd ) − 2ω(σ + µ)eστd = 0 (29)

K(σ 2 + 2µσ − ω 2 ) cos(ωτd ) + 2Kω(µ + σ) sin(ωτd ) + (σ 2 + 2µσ − ω 2 + ω◦2 )eστd = 0 (30)


The stability of the system depends on the value of the parameter σ. The system is asymptotically stable for σ < 0
and unstable for σ > 0. To determine the boundaries of linear stability and instability we set σ equals to zero in
Equations (29) and (30), and with simple manipulation, we find
p
4ν 2 + λ2 (λ2 + 4ν 2 − 1)2
K(λ) = − (31)
λ(λ2 + 4ν 2 )
· µ ¶ ¸
1 −1 2ν
γ(λ) = tan + jπ j = 0, 1, 2, ... (32)
2πλ λ(λ2 + 4ν 2 − 1)
where λ = ω/ω◦ , γ = τd /T , and ν = µ/ω◦ . Varying λ and solving these equations for K and γ, we determine the
stability boundaries. Figure 6 shows the stable and the unstable regions as predicted by the linear theory. Any gain-
delay combination that lies inside the unshaded areas of Figure 6 leads to a stable response. It is worth mentioning
that there are infinite number of stability pockets, which decrease in size as the time delay of the controller τd
increases.

To determine the magnitude σ of the damping resulting from each gain-delay combination, we vary τd and K in
Equation (29) and (30) and calculate σ. Figure 7 shows a contour plot of the value ofσ in the first pocket of stability.

Figure 6: A stability plot of the delayed position-feedback controller for a relative damping ν = 0.0033. The shaded areas represent the pockets
of stability.
Figure 7: A contour plot of the damping as a function of the gain K and the delay τd , where τd is given in terms of the natural period of the
uncontrolled system. The darker the areas are the higher the damping is.

5. SIMULATIONS
We carried out several numerical simulations on the full model of the container crane. The following values are used
for the model parameters in the simulation: R = 2.5 m, d = 2.825 m, w = 1.4125 m, and m = 50000 kg. For the IS
controller we use n = 1(0.5Tc < τs < 1.5Tc ) and S = 50 m, and for the delayed-position feedback controller we use
K = 0.4 and τd = 0.28 T .

Shaped acceleration profiles based on the NSP, the LDP, and the NDP frequency approximations are generated
then applied to the full 4-bar mechanism model, Figure 8. Results show that the NSP acceleration profile, will not
reduce the residual oscillations. On the contrary, this shaped profile will amplify residual oscillations to magnitudes
that are even larger than the transient oscillations. The LDP and NDP models did reduce the residual oscillations
to significantly small magnitudes. However, the figure shows that the residual oscillations resulting from the NDP
acceleration profile has a much better performance than the LDP profile.

In Figure 9, we demonstrate the sensitivity of open-loop input-shaping controllers to changes in system parameters.
We change the cable length and apply the same NDP acceleration profile shown in Figure 8a to the full model. Sim-

1.5 4
NSP NSP
LDP LDP
NDP NDP
3
1

0.5
1
Acceleration[m/s2]

Sway[m]

0 0

−1
−0.5

−2

−1
−3

−1.5 −4
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40

Time [sec] Time [sec]

(a) Shaped operator commands. (b) Time series of the sway .

Figure 8: Sway response of the full of the container crane to shaped operator commands. Results are obtained for L = 17.5m.
4
16.5 m
17.5 m
18.5 m
3

Sway[m]
0

−1

−2

−3

−4
0 5 10 15 20 25 30 35 40

Time [sec]

Figure 9: Sway response of the full model of the container crane to the shaped operator commands shown in Figure (13-a)

ulations show that a change of 1.0 m around the design value of the cable length amplifies the residual oscillations
five orders of magnitude.

To overcome the problem of the open-loop input-shaping controller sensitivity, and to eliminate the residual oscil-
lations, we apply a delayed-position feedback controller at the end of the transfer maneuver. To demonstrate the
performance, an NDP acceleration profile based on a 17.5 m cable is applied to a 16.5 m cable. Figure 10 shows
that an input-shaping controller based on the NDP frequency approximation results in a residual sway of approxi-
mately 1.25 m. On the other hand a combination of input-shaping and and delayed-position feedback controller (ED)
suppresses the residual sway to a magnitude less than 0.05 m within 4.5 s of the end of the transfer maneuver.

6. CONCLUDING REMARKS
The bang-off-bang input-shaping controller based on a simple pendulum model fall short of satisfying the goal of
eliminating residual oscillations on quay-side container cranes. Operating at a constant cable length and using a
nonlinear frequency approximation, the residual oscillations are many orders of magnitude larger than those when
the same shaped acceleration profiles are applied to the simple pendulum model.

To achieve satisfactory performance, input-shaping controllers should be based on accurate mathematical models.
Our results show that a constrained double pendulum model based on a four-bar mechanism model of a container
crane is currently the best existing approximate model for quay-side container cranes. For predefined system
parameters, the input-shaping acceleration profiles based on a nonlinear frequency approximation of this model are
capable of reducing the residual oscillations to magnitudes less than 0.01 m.

1.5 4
IS IS
IS + ED IS + ED

3
1

0.5
1
Acceleration[m/s2]

Sway[m]

0 0

−1
−0.5

−2

−1
−3

−1.5 −4
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40

Time [sec] Time [sec]

(a) Input acceleration profile. (b) Time series of the sway .

Figure 10: Sway response of the full model of the container crane. Results are obtained for L = 16.5 m.
A delayed-position feedback controller is used to minimize the sensitivity of the input-shaping controller to variations
View publication stats

in system parameters and to improve its robustness. The addition of such feedback controller will account for system
delays, external disturbances, and will compensate for uncertainties in the system. The delayed-position feedback
controller is capable of reducing the residual oscillations of the payload to less than 0.05 m within 5 s of the end of
the transfer maneuver, while maintaining the target trolley position.

A number of factors were considered behind the choice of applying a feedback controller only at the end of the
transfer maneuver. A significant factor is reducing the power required to perform a transfer maneuver with minimal
residual oscillations. This stems from the fact that many feedback control systems require input accelerations that
are beyond the normal operating accelerations, which may overload the trolley motors. However, when applied at
the end of the transfer maneuver performed using an input-shaping control system, feedback controllers require
lower accelerations. This is due to the fact that the input-shaping controller produces reduced residual oscillations.

Another important factor is improving both the predictability and smoothness of the trolley motion. Feedback con-
trollers produce continuously changing acceleration profiles, which affects the crane operator performance and
comfort. Therefore, applying an efficient feedback controller, such as the delayed-position feedback, at the end
of the transfer maneuver for a very short period of time eliminates excessive trolley motion. Thus maintaining
comfortable working conditions for the crane operator.

References
[1] Masoud, Z., and Nayfeh, A. H., Sway Reduction on Container Cranes Using Delayed Feedback Controller, in
Proceedings of the AIAA, 2002.
[2] Hubbel J. T., Koch, B., and McCormic, D., Modern Crane Control Enhancements, in Ports 92, Seattle, WA,
1992, 757-767.
[3] Alsop, C. F., Forster, G. A., and Holmes, F. R., Ore Unloader Automation - A Feasibility study, in Proc. of IFAC
Workshop on Systems Engineering for Control Systems, 1965, Tokyo, Japan, 295-305.
[4] Alzinger, E., and Brozovic, V., Automation and Control System for Grab Cranes, Brown Boveri Review, 7, 1983,
351-356.
[5] Jones, J. F., and Petterson, B. J., Oscillation Damped Movement of Suspended Objects, in Proc. of the IEEE
International Conference, Seattle, WA, 1995, 956-962.
[6] Dadone, P. , and Vanlandinham, H. F., Load Transfer Control for a Gantry Crane with Arbitrary Delay Con-
straints, Journal of Vibration and Control, 7, 2001, 135-158.
[7] Pyragas, K., Continuous Control of Chaos by Self-Controlling Feedback, in Phys. Lett., A 170, 1992, 421-428.
[8] Gregory, D. V., and Rajarshi R., Chotic Communication Using Time-Delayed Optical Systems , in International
Journal of Bifuractions and Chaos., 9, 1999, 2129-2156.
[9] Pacora L. M., and Carroll T. L., Synchronization in Chaotic Systems, in Physics Review Letter., 64, 1990,
821-824.
[10] Xu. J., and Chung K. W., Effects of Time Delayed Psoition Feedback on a Van Der Pol-Duffing Oscillator, in
Physica., D 180, 2003, 17-39.
[11] Nayfeh, A. H., Ching C.-M., and Pratt J., Perturbation Methods in Nonlinear Dynamics-Applications to Machin-
ing Dynamics, in Journal of Manufacturing Science and Engineering., 119, 1997, 485-493.
[12] Cheng, C. C., and Chen C. Y., Controller Design of an Overhead Crane System with Uncertainties, in Control
Engineering Practise, 4, 645-653.
[13] Henry, R., Masoud, Z., Nayfeh, A. and Mook, D., Cargo Pendulation Reduction on Ship-Mounted Cranes via
Boom-Luff and Slew Angles Actuation, Journal of Vibration and Control, 7, 2001, 1253-1264.
[14] Nayfeh, A. H., Introduction to Perturbation Techniques, 1981, Wiley, New York.

You might also like