1 s1767 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Applied Mathematical Modelling 108 (2022) 807–824

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Neural fractional-order control of telescopic truck cranes


Le Anh Tuan a,∗, Le Van Duong b
a
Vietnam Maritime University, Vietnam
b
Le Quy Don Technical University, Vietnam

a r t i c l e i n f o a b s t r a c t

Article history: A telescopic boom crane is a strictly underactuated system in which three actuators track
Received 8 January 2022 five motions. The boom luffing and extension cause variations in the shape and structure
Revised 1 April 2022
of cranes. The dynamic reaction between payload and boom through a flexible rope and
Accepted 4 April 2022
viscoelastic cylinders induces significant vibrations. Additionally, a crane contains many
Available online 9 April 2022
uncertain parameters, suffers wind disturbances and sometime encounters failure of actu-
Keywords: ators. On the basis of a complex model with five fully nonlinear differential equations, we
Fractional-order control construct an intelligent robust control system for such cranes, which integrates the benefits
Mittag-Leffler stability of advanced techniques. Fractional-order control combined with robust control produces a
Neural network control controller holding flexible derivative orders. A multilayer perceptron network provides a
Robust adaptive control learning feature to the cranes, in which an adaptive mechanism estimates one equivalent
Sliding mode control component represented for individual influences. Simulation for a practical crane shows
Telescopic cranes the effectiveness of the proposed control system.
© 2022 Elsevier Inc. All rights reserved.

1. Introduction

Telescopic truck cranes work frequently at construction sites to lift and transfer loads to desired positions. A boom crane
mounted on a truck or a carrier is normally used to enable mobility. As its name implies, a crane has an expandable boom
with telescopic tubes conducting relative motions. In pick-and-carry operation, a crane changes lift height and horizontal
reach by luffing and extending the telescopic boom. In old versions, a telescopic boom uses many hydraulic cylinders for
longitudinal motions, and its segments must be pinned while carrying payload. Modern truck cranes allow boom expansion
when hoisting the load and luffing the boom. They use only one hydraulic cylinder together with rope transmission for
telescoping the boom. The boom of a heavy-duty crane holds massive volume. When luffing the boom, the centre translation
of weight mass and high moment of inertia easily leads to instability and difficult control of motion. Modern truck cranes
lighten the boom weight through shape optimisation and material solution. They equip advanced control technologies that
allow simultaneous combination of motions. Increasing the productivity requires fast motion in modern cranes. This causes
imprecise motions of crane mechanisms, large sway of cargo and boom swing, which may lead to danger and even an
accident at the operation site. In this work, we develop a state-of-the-art control system for such cranes, in which the cargo
is fast-tracked to precise destinations without much load pendulation and crane vibrations.
A telescopic boom crane (TBC) is a complex mechanical system. It includes hundreds of rotating and translating compo-
nents connected by rigid and elastic links. Geometrical configuration and kinematic constraints make a crane be a strictly


Corresponding author.
E-mail address: tuanla.ck@vimaru.edu.vn (L.A. Tuan).

https://doi.org/10.1016/j.apm.2022.04.006
0307-904X/© 2022 Elsevier Inc. All rights reserved.
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

nonlinear system. The forces and torques at motors and actuators, together with wind loads, induce vibrations of crane
components.
A TBC is also a strictly underactuated system [1–4]. To increase the productivity, many actuators run at the same time. For
instance, three actuators produce five motions: rotation of hoist drum, boom swing, boom extension, payload-lifting motion
and payload swing. The crane holds elastic links, such as handling cables and two hydraulic cylinders that contribute to the
increment in vibrations, especially load oscillation along the cables, load swing and boom swing. Lack of actuators, tracking
the elastic multibody system while suppressing many unwanted vibrations, the controller meets many challenges.
Furthermore, a TBC is a shape-varying system. It holds many uncertain parameters, suffers wind disturbance and some-
time faces failure of actuators. The motions of main components, such as boom expansion, boom luffing, longitudinal oscilla-
tion and payload swing, cause variations in the shape, structure and dimensions of a crane. In addition to facing such struc-
tural uncertainty, a TBC contains parametric (unstructured) uncertainty. The volume and weight of payload are changeable
for each duty-cycle operation. Working at various environments leads to uncertainty in machine friction. In this study, we
regard load mass and damping coefficients as uncertain parameters. Working outdoor, a TBC is also unavoidably influenced
by random wind. The difficulty increases when faults occur at hoisting motors and/or at hydraulic cylinders. Controlling a
TBC holding both structural and unstructured uncertainties, subjected to unknown wind and motor failure, has never been
easy.
Various versions of boom cranes are operated in industry and transportation. In accordance with lower structure, they
comprise fixed-base cranes [5], crawler cranes [6], truck-mounted cranes [7] and shipboard cranes [8,9]. In accordance with
boom frame, they are composed of fixed-length boom cranes [10–14], TBCs [15–23], knuckle boom cranes [24] and lattice
boom cranes [6]. In comparison with other boom cranes, TBCs occupy a small number of related studies in modelling and
control. The reason may be their complex configuration, especially the shape variation of boom. A TBC controller conducts
many goals concurrently. It not only tracks and regulates control outputs but also keeps the overall crane stable despite the
transposition of mass centre and the shape variation of the crane.
The modelling of TBCs consists of static model [15,24–26] and crane dynamics [7,16−20,23]. While static problems dis-
cuss balance, static stability [25] and deformation of the crane structure [15,26], the crane dynamics builds equations of
motion, finds the solutions (crane responses) via analytical methods or numerical analysis and illuminates the physical
features and dynamic behaviour of cranes. With model-based control approaches, the control formation is more precise
if the dynamic model is close to practical operation. Owing to the complexity of a TBC, a dynamic model representing
all physical features of a crane is impossible. In this study, we attempt to provide a dynamic model that considers many
structural factors, together with real influences on cranes during their operation. We regard a TBC as an elastic multibody
system suffering both structural and unstructured uncertainties. Payload ropes, luffing cylinder and telescopic cylinder are
considered elastic links containing internal frictions. We explore a complex operation scenario in which three actuators
comprising hoisting motor, boom-lifting cylinder and boom-expanding cylinder are simultaneously activated to produce five
motions: longitudinal oscillation of payload, payload swing, boom swing, boom extension and rope-rolling motion. The dy-
namic model comprises a set of five fully nonlinear differential equations. Such a model describing almost all of dynamic
behaviours improves the precision of control system design.
Dynamic modelling effectively supports the design of control systems by using model-based control approaches [8–14].
Control of TBCs is harder than that of fixed-length boom, gantry and overhead cranes. The main reasons lie in the insta-
bility of the cranes due to swinging, retracting or expanding the heavy boom. In literature, the publications on the control
of fixed-length boom cranes are significant in comparison with those on the control of TBCs [12,21,22]. Generally, a control
system works on the basis of the feedback principle. However, feedforward control is an exception. Time-delay technique
and command shaping are open-loop controls applied to TBCs to suppress the residual vibration of the boom and payload
swing [21]. They do not need sensors for measuring feedback signals. Such control systems are compact and thus hold price
reduction and economic benefits. Nevertheless, the controllers are sensitive to the variation in payload weight, boom length
and wire rope length. A large variation in these components may make the crane unstable. Feedback controllers give a
command or decision on the basis of feedback information. Although PID controllers [10,22] are classical, they are effective
for boom cranes. Studies [10,22] have verified the promising application of PID control to industrial boom cranes. Indeed,
PID holds a simple structure and is easy to implement; however, it is not robust. Its control gains are sensitive to wind
disturbance and uncertain payload mass. With no way to adaptively tune control gains, the cranes become easily unstable
when meeting large uncertainties. Pole placement combined with LQR [11] is another control solution. This way is feasible
for linear systems, but truck cranes are strictly nonlinear systems. Using a linearised crane model leads to the imprecision of
control formation. Without the need to linearise the dynamic model, the nonlinear feedback technique [14] utilises a nonlin-
ear crane model to design controllers. Its control scheme includes two components: one is to stabilise the outputs, and the
other is to suppress the crane nonlinearities. In fact, pole placement [11], PID [10,22] and feedback linearisation [14] are not
useful when cranes work at uncertain sites and suffer random wind load. These controllers may not keep the consistency of
output responses when cranes face a wide variation in their structure and parameters, as well as strong disturbances. Sliding
mode control (SMC) [9,12] is the other approach that tends to the system robustness. This technique has been successfully
applied to two versions of boom cranes: fixed-length boom crane [12] and knuckle boom crane [9]; one was mounted on
a truck [12], and the other was mounted on a ship [9]. While feedback linearisation [14] needs to provide as much model
knowledge as possible, SMC does not require high accuracy of the crane model [27,28]. It still well tracks the outputs to
destinations and remains the output robustness despite uncertainties, winds [12] or sea wave actions [9]. Although SMC is

808
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

a robust control approach, it is not intelligent. An SMC controller does not change its structure to adapt to the variation
in loads, disturbances and other uncertainties. It persists in stiffly keeping the outputs consistent without adaptation and
flexibility. Other approaches, such as repetitive learning control [13] and self-tuning control [9], are integrated to improve
the adaptability of control systems. The controllers either automatically adjust their gains or estimate uncertainties with
the support of adaptive mechanisms. Recently, the application of fractional calculus to engineering problems has attracted
many researchers [29–31]. In control theory and application, fractional derivative is normally combined with various control
techniques, such as optimal control [32] and robust control [33,34], to establish a control approach, that is, the so-called
fractional-order control (FOC). Tending to a control system for TBCs holding both robust and adaptive features, we combine
three control techniques; thus, the control system achieves the benefits of these control methods. The control core com-
prises an SMC-like method and FOC. It contains the sign function for tracking errors to maintain the output’s robustness.
It is formed with the variable fractional derivative of feedback signals instead of first-order derivative as usual. Meanwhile,
an adaptive mechanism using a neural network is constructed to address uncertainties, faults and unknown winds. Lying
on the feedback loop of the control system, the adaptive mechanism utilises crane outputs and desired inputs to achieve
approximation of parametric variation, structural uncertainties, winds and faults. The core of this mechanism is a neural
network and a fractional adaptation algorithm for training and estimating the unknown equivalence. The proposed control
system holds two key features.

(i) The feedback signals comprise signals q measured by sensors and fractional derivative signals Dα q. In a traditional
control system, it is a pair of (q, q˙ ).
(ii) The core of the controller produces the crane’s robustness, and the estimator realises crane adaptation. Both control
and estimation algorithms hold fractional structures.

This work contributes the following state of the art:

(i) While typical control structures hold integer orders of derivative, FOC applied to both controller and adaptive mech-
anism generates adjustable fractional orders (FOs). They play a role as control parameters and adaption gains. This
leads to a flexible control structure that is convenient for optimal and adaptive control.
(ii) Together with strict underactuation, a TBC is regarded as an elastic multibody system holding flexible links, uncertain
structure, shape variation and time-varying parameters. Model-based control with high model accuracy leads to an
excellent control system.
(iii) The control system well tracks the crane outputs, stabilises the entire crane despite the variation in the crane shape
and dimension and works adaptively and flexibly with the presence of uncertain parameters, uncertain structure,
random wind and faults of actuators. A neural network and estimation algorithms improve the learning skills of the
crane and make the crane intelligently adapt to all unfavourable impacts. Such a controller has not been applied to
TBCs, until recently.

The research contents are organised as follows. Section 2 provides a mathematical model of a TBC considering a complex
operation, in which running three actuators results in five output motions. The crane model is regarded as an elastic multi-
body nonlinear system in the presence of structural and unstructured variations, winds and actuator faults. Section 3 anal-
yses and synthesises a control system mainly on the basis of three approaches: FOC, robust control and adaptive control
with neural networks. Aside from tracking crane outputs, the adaptive mechanism conducts multiple goals simultaneously:
compensating for random winds and detecting and estimating crane uncertainties and faults. Section 4 investigates the
effectiveness of the proposed approach by applying it to an industrial TBC, namely, Liebherr LTF 1060. An analysis and
a comparison with various control approaches are also included. Lastly, conclusions and further research are indicated in
Section 5.

2. TBC dynamics

2.1. System description

We establish a dynamic model of a TBC, whose diagram is depicted in Fig. 1. A crane can combine many motions for
improving the productivity and speeding up load transportation. In this operation scenario, payload mp is picked up and
carried by simultaneously activating three actuators: hoisting payload with motor moment Mh , lifting boom with cylinder
force Fl and expanding boom with telescopic cylinder force Fe . Correspondingly, five motions are created: rotation φ of
hoisting drum Jh , boom swing ψ , boom extension q, longitudinal motion s of payload and payload swing θ . The boom
comprises two segments of rigid bodies with physical parameters lt , Jt , mt and lb , Jb , mb . The load-carrying rope is regarded
as an elastic-massless string represented by kp and bp . The elasticity ke and kl of the two cylinders contributes to axial
payload oscillation, boom swing and angular vibration of the hoisting drum. Meanwhile, the viscosities be and bl reduce
these vibrations. The other parameters are lAB , lCD , l(t), c=GA, d=KA, h=GE, k=KM, n=IA, m=ML, ho =AO, r=JA, λ = IAB  and

γ = HAK .

809
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 1. Diagram of upper structure of a telescopic boom crane.

2.2. Equations of motion

Through running three actuators Ua =[Mh Fl Fe ]T , five motions with respect to five generalised coordinates q=[qi ]T =[φ ψ
q s θ ]T are produced. The physical nature of a TBC is governed by five second-order differential equations of motion:

Jh φ̈ + b p R2h φ˙ + b p Rh as˙ + b p Rh q˙ + k p Rh [Rh (φ − φ0 ) + a(s − s0 ) + q − q0 ] + Rh m p g/a = Mh + 1 φ1 , (1)

{Jt + Jb + 0.25mb lb2 + m p [0.5(lt + lb ) + q]2 + m p k2 + mt (0.5lb + q )2 }ψ̈ + m p kq̈


+m p {k sin(θ − ψ ) − [0.5(lt + lb ) + q] cos(θ − ψ )}s̈ + m p s{k cos(θ − ψ ) + [0.5(lt + lb ) + q] sin(θ − ψ )}θ̈
+{[m p lt + (m p + mt )lb + (2mt + m p )q]q˙ + bl n2 d2 sin (γ − λ + ψ )/ll2 (ψ )}ψ˙
2

+m p {2s˙ θ˙ [0.5(lt + lb ) + q] − ksθ˙ (θ˙ − ψ˙ )} sin(θ − ψ ) + m p {sθ˙ 2 [0.5(lt + lb ) + q] + ks˙ (2θ˙ − ψ˙ )} cos(θ − ψ )
 
+{m p [0.5(lt + lb ) + q] + 0.5mb lb + mt (0.5lb + q )}g cos ψ − m p k sψ˙ θ˙ sin(θ − ψ ) − ψ˙ s˙ cos(θ − ψ ) − g sin ψ
+kl [ lls + ll (ψ ) − ll (ψ0 )]nd sin(γ − λ + ψ )/ll (ψ ) = Fl cos ψ − wB sin ψ + 2 φ2 , (2)

m p kψ̈ + (m p + mt )q̈ + m p sin(θ − ψ )s̈ + m p s cos(θ − ψ )θ̈ + (b p + be )q˙ + b p Rh φ˙ + b p as˙


− {0.5m p lt + 0.5(m p + mt )lb + (m p + mt )q}ψ˙ 2 − m p sθ˙ 2 sin(θ − ψ ) + 2m p s˙ θ˙ cos(θ − ψ ) + ke ( les + q − q0 )
+ k p [Rh (φ − φ0 ) + a(s − s0 ) + q − q0 ] + m p g/a + (m p + mt )g sin ψ = Fe + wc cos ψ + 3 φ3 , (3)

m p {k sin(θ − ψ ) − [0.5(lt + lb ) + q] cos(θ − ψ )}ψ̈ + m p sin(θ − ψ )q̈ + m p s̈ + b p a2 s˙ + k p a2 (s − s0 )


+b p a(Rh φ˙ + q˙ ) + k p a[Rh (φ − φ0 ) + q − q0 ] − m p (kψ˙ 2 + 2q˙ ψ˙ ) cos(θ − ψ )
−m p ψ˙ 2 [0.5(lt + lb ) + q] sin(θ − ψ ) − m p sθ˙ 2 + 2m p gsin (θ /2 ) = 0,
2
(4)

m p s{[0.5(lt + lb ) + q] sin(θ − ψ ) + k cos(θ − ψ )}ψ̈ + m p s cos(θ − ψ )q̈ + m p s2 θ̈ + 2m p ss˙ θ˙


+m p sψ˙ (2q˙ + kψ˙ ) sin(θ − ψ ) − m p sψ˙ 2 [0.5(lt + lb ) + q] cos(θ − ψ ) + m p gs sin θ = w p s cos θ . (5)

810
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Eqs. (1)–(3) are simplified in matrix form:

Maa (q )q̈a + Mau (q )q̈u + Ba (q, q˙ )q˙ a + Ga (q ) + Fa (q, q˙ ) = Jua (q )Ua + Jwa (q )Wa + (t − t f )(q, q˙ , Ua ) (6)

describing the features of actuated outputs qa =[φ ψ q]T . Eqs. (4) and (5), also in matrix form

Mua (q )q̈a + Muu (q )q̈u + Bu (q, q˙ )q˙ u + Gu (q ) + Fu (q, q˙ ) = Jwu (q )Wu (7)
   
m11 0 0 0 0
represent unactuated states qu =[s θ ]T . Here, Maa = 0 m22 m23 ,Mau = m24 m25 = MTua and Muu =
0 m32 m33 m34 m35
 
m44 0
are inertial matrices, where Maa and Muu are symmetric positive definite; Ba = diag(b1 , b2 , b3 ) and
0 m55
Bu = diag(b4 , b5 ) are damping matrices; Ga = [g1 g2 g3 ]T and Gu = [g4 g5 ]T are gravitational vectors; Fa =
[ f1 f2 f3 ]T and Fu = [ f4 f5 ]T are vectors of nonlinearities; Jua = diag(1, cos ψ , 1 ), Jwa = diag(0, − sin ψ , cos ψ ) and
Jwu = diag(0, s cos θ ) are Jacobian matrices; and Wa = [0 wB wC ]T and Wu = [0 w p ]T are two subvectors of wind load
W. The elements of the above-mentioned matrices and vectors are provided in the Appendix.
(t − t f )(q, q˙ , Ua ) denotes the faults of actuators, where  = diag(1 , 2 , 3 ) are the gains of faults, tf is the time
range of fault occurrence, and  = [ 1 2
T
3 ] indicates the fault functions given by

0 t < tf
(t − t f ) = (8)
i
1 − e−ai (t−t f ) t ≥ tf ,

Eq. (6) shows the direct actions of control actuators Ua and their faults  in actuated states qa , while no explicit reaction
between Ua and unactuated states qu is observed in Eq. (7). In fact, control inputs Ua rely on the kinematics and geometric
constraints of the motions to act on qu . To make this obvious, we withdraw q̈u from Eq. (7) and insert it into Eq. (6), then
we obtain

Meq (q )q̈a + Ba (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) = Jua (q )Ua + Weq + (t − t f )(q, q˙ , Ua ) (9)

where

Meq (q ) = Maa − Mau M−1


uu Mua , (10)

Beq (q, q˙ ) = −Mau M−1


uu Bu , (11)

Geq (q ) = Ga − Mau M−1


uu Gu , (12)

Feq (q, q˙ ) = Fa − Mau M−1


uu Fu , (13)

And:

Weq = Jwa Wa − Mau M−1


uu Jwu Wu . (14)

The reduced-order system (9) is equivalent to the original system (6) and (7). The physical features of unactuated outputs
described in dynamics (7) are integrated into actuated dynamics (9).
We consider three time-varying parameters p = [ m p bp bl ]T of the crane corresponding to uncertaintiesp ˜=
[ m˜p ˜
bp ˜ T
bl ] . In this case, the equivalent dynamics (9) is enhanced as:

Meq (q )q̈a + Ba (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) = Jua (q )Ua + H(q̈, q˙ , q, p
˜ , , W ) (15)

where:

H(q̈, q˙ , q, p
˜ , , W ) = Weq + W
˜ eq − (M
˜ eq q̈a + B
˜ a q˙ a + B ˜ eq + F˜ eq ) + (t − t f )(q, q˙ , Ua )
˜ eq q˙ u + G (16)

is an equivalence of faults , unknown wind Weq and its uncertainty W ˜ eq and uncertainties M ˜ eq (m
˜ p ), B
˜ a (m
˜ p , b˜ p , b˜ l ),
Beq (m
˜ ˜ p , b p ), Geq (m
˜ ˜ ˜ p ), Feq (m
˜ ˜ p , b p ), Weq (m
˜ ˜ ˜ p ) parameterised in terms of parametric variation p
˜ . In accordance with mean
value theorem, the equivalence (16) is uniformly upper bounded by:

H(q̈, q˙ , q, p˜ , W ) ≤ H0 + H1 q + H2 q2 (17)

With H0 , H1 and H2 as positive coefficients.

811
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

3. Control system design

In a traditional control system, the outputs are displacements q, while the feedback signals include q and their velocities
q˙ . Truck cranes are equipped with a wireless gyro-sensor to measure the angular velocity of load and encoders to measure
the boom angle, translation of telescopic cylinders and rotation of the upper frame. This section provides two main parts for
a crane control system. The first part is a robust control core. Using three actuators Ua =[Mh Fl Fe ]T , the control core tracks
three actuated outputs qa =[φ ψ q]T to destinations qad =[φ d ψ d qd ]T and quickly stabilises two unactuated outputs qu =[s
θ ]T around equilibrium qu =[sd 0]T with φ d =(s0 −sd )a/Rh . The second part is a neural network-based adaptive mechanism for
estimating parametric uncertainties p ˜ and unknown disturbances W (or Weq ) and compensating for the faults of actuators.
Fractional calculus is applied to both control core and adaptive mechanism. We thoroughly use the Caputo derivative with
0 < α ≤ 1 in this study.

3.1. Fractional control core

Notably, Meq (q) is a positive-definite matrix. We acquire the double-time derivative of actuated outputs qa from equiva-
lent dynamics (15) as:
 
eq (q ) Jua (q )Ua + H (q̈, q
q̈a = M−1 ˜ , , W ) − Ba (q, q˙ )q˙ a − Beq (q, q˙ )q˙ u − Geq (q ) − Feq (q, q˙ )
˙ , q, p (18)
Let:

V = D2α qa = D2(α −1) q̈a ∈ R3 (19)


be a new control input. Here, D2α q a denotes the fractional derivative of qa having order α ∈(0,1]. If introducing an interme-
diate input

V = D2α qad − 2λDα ea − λT λea , (20)


we will hold a stable FO equation of actuated states qa

D 2 α e a + 2 λD α e a + λT λe a = 0 (21)
that assures twice Mittag–Leffler convergence of ea (or qa converges to qad )

ea = ea0 E−λt (22)


for every positive diagonal matrix λ = diag(λ1 , λ2 , λ3 ). Here, E(•) is a Mittag–Leffler function that converges faster than an
exponential function. For tracking qa only, we can use a control core deduced from Eqs. (18)–(20) as follows:
 
B (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) − H(q̈, q˙ , q, p
˜ , , W )
Ua = J−1 (q ) a (23)
ua −Meq (q )D2(1−α ) 2λDα ea + λT λea

where ea =qa −qad and eu =qu −qud are locally tracking errors. For achieving the goal of stabilising unactuated states qu , the
control core (23) is enhanced into:
 
B (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) − H(q̈, q˙ , q, p
˜ , , W )
Ua = J−1 (q ) a (24)
ua
−Meq (q )D2(1−α ) 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne
 T
ρ1 0 ρ3
whereη = diag(η1 , η2 , η3 ) is a positive-definite matrix, ρ = is a matrix of positive gains for adjusting the
0 ρ2 0
influence of Ua on qu , and

e = [e i ]T = D α e a + λ e a + ρ e u (25)
is a globally tracking error in the form of fractional derivative.

Theorem 1. Despite the faults  of three actuators, unknown wind Weq , wind uncertainty W ˜ eq and parametric uncertain-
˜ , the robust control core (24) with manifold (25) tracks the outputs q=[qa qu ]T of a TBC governed by dynamics (6) and
ties p
(7) to destinations qd =[qad qud ]T asymptotically.

Proof. The control core (24) comprises five terms playing individual roles. Term Ba (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ )
is to suppress the nonlinearities of a crane, and term H(q̈, q˙ , q, p
˜ , , W ) is for compensating for faults , wind W and para-
metric uncertainties p˜ . Meanwhile, term (2λDα ea + λT λea ) converges actuated states, (ρDα eu + λρeu ) stabilises unactuated
states, and ηsgne is to maintain the robustness of qa and qu at desired positions located in regime e. We investigate the
crane stability by considering a positive Lyapunov function.
1 T
V = e e>0 (26)
2

812
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

whose fractional derivative is:


1 α T
Dα V = D (e e ) (27)
2
which yields:
Dα V ≤ eT Dα e (28)
in accordance with Lemma 4 of [35]. Meanwhile, crane dynamics (15) subject to control inputs (24) induces a closed-loop
system
D2α ea + 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne = 0 (29)
or equivalently:
Dα e + λe + ρDα eu + λρ eu + ηsgne = 0 (30)
that leads the Lyapunov derivative (28) to
Dα V ≤ −eT λe + ρDα eu + λρ eu + ηsgne
≤ −eT λe − eT ηsgne − eT ρDα eu + λρ eu (31)

If eT (Dα ρ e u + λρeu ) ≥ 0, then the fractional derivative (31) of the Lyapunov function becomes:
3 3
Dα V ≤ − λi e2i − ηi |ei |
i=1 i=1
3 3
≤ −λmin e2i − ηmin |ei | = −λmin e2 − ηmin e
i=1 i=1

≤ −λmin e2 (32)


where λmin = min (λi ) > 0, and ηmin = min (ηi ) > 0.
The Lyapunov function (26) V = 0.5eT e = 0.5e2 and its derivative (32) satisfy Theorem 6.2 of [36]. That is, equilibrium
e=0 is asymptotically stable. The zero convergence of e (25) yields
D α e a + λe a + ρ e u = 0 (33)
and reduces the closed-loop Eq. (30) to
ρDα eu + λρ eu = 0 (34)
We consider the auxiliary Lyapunov function
1 1
Va = (ρeu )T ρeu = eTu ρT ρeu > 0 (35)
2 2
ρ12 + ρ32
0
with ρT ρ = [ ] being a positive-definite matrix. Remarkably, eu ∈ R2 and ρeu ∈ R3 have the same sign, given
0 ρ22
that ρ is a matrix of positive gains.
The fractional derivative of auxiliary Lyapunov (35) with the application of Lemma 4 of [35] is in the form of
1 α 
Dα Va = D (ρeu )T ρeu ≤ (ρeu )T Dα (ρeu ) (36)
2
Combining it with Eq. (34) makes the Lyapunov derivative (36) be
Dα Va ≤ −(ρeu )T λρ eu = −eTu ρT λρ eu ≤ 0 (37)
+ 0ρ2λ ρ32 λ3
considering that ρT λρ = [ 1 1] is a positive-definite matrix.
0 ρ22 λ2
The Lyapunov function (35) is bounded by
1
α1 eu 2 ≤ Va = eTu ρT ρeu ≤ α2 eu 2 (38)
2
which yields
α1 eu  ≤ Va ≤ α2 eu  (39)
where α 1 and α 2 are arbitrary positive constants satisfying α1 ≤ 0.5 min(ρ12 λ1+ ρ32 λ3 , ρ22 λ2 ) and α2 ≥ 0.5 max(ρ12 λ1
+
ρ32 λ3 , ρ22 λ2 ). In expression (39), Va ≥ α 1 eu  means that Va is positive definite, and Va ≤ α 2 eu  implies that it is de-
crescent.

813
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

The Lyapunov derivative (46) can be rewritten as


Dα Va ≤ −eTu ρT λρ eu ≤ −α3 eu 2 (40)
or equivalently
Dα Va ≤ −α3 eu  (41)
with α 3 being an arbitrary positive constant satisfying α3 ≤ min(ρ12 λ1 + ρ32 λ3 , ρ22 λ2 ).
The Lyapunov candidate (39) and in-
equality (41) fit with Theorem 6.2 of [36]. That is, eu presents Mittag–Leffler stability, eu (t ) ≤ eu (0 )E−α3 t , or qu ap-
proaches qud asymptotically. Near the origin, the Mittag–Leffler stability of the FO system shows faster convergence com-
pared with the exponential stability of the integer-order systems [36].

Remark 1. The discontinuous term ηsgn(e) of controller (24) may cause chattering and even instability at large values of η.
These can be prevented via several ways, such as to replace the signum function with a boundary layer, a sigmoid function
or state-dependent gains.

3.2. Neural network-based adaptation

Remarkably, the control core (33) cannot tackle well the case that the knowledge of motor faults and disturbances is not
provided. Although it can maintain the system’s robustness when facing uncertainties and bounded disturbances, it is not
flexible and intelligent. The control structure is not changeable with fixed gains. Therefore, we propose a control law with
a flexible structure and an adaptive mechanism using neural networks for concurrently compensating for faults of actuators
and unknown wind load and estimating the variations of crane parameters.
We simplify the control core (24) in the form of
 
ua (q ) L (q̈, q
Ua = J−1 ˜ , , W ) − Meq (q )D2(1−α ) 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne
˙ , q, p (42)
whose adaptive form is
 
ua (q ) L (q̈, q
Ua = J−1 ˆ ˜, 
˙ , q, p ˆ ) − Meq (q )D2(1−α ) 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne
ˆ ,W (43)

where Lˆ (q̈, q˙ , q, p
˜,  ˆ ) is an estimation of equivalent component
ˆ ,W

L(q̈, q˙ , q, p
˜ , , W ) = Ba (q, q˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) − H(q̈, q˙ , q, p
˜ , , W ) (44)
bounded by

L(q̈, q˙ , q, p˜ , , W ) ≤ L0 + L1 q + L2 q2 (45)


Hence, instead of individually approximating nonlinearities Ba (q, q˙ )q˙ a ,Beq (q, q˙ )q˙ u ,Geq (q),Feq (q, q˙ ),faults , parametric
uncertainties p˜ and wind disturbances W, a neural mechanism lying on a feedback loop will approximate only one unique
component Lˆ (q̈, q˙ , q, p
˜,  ˆ ). This contributes to the reduction in burden computation and estimation. As an upgrade of
ˆ ,W
Theorem 1, the following theorem provides such a flexible controller and an estimator to adapt to the presence of nonlin-
earities, parametric uncertainties, unknown wind and actuator faults.

Theorem 2. We consider an adaptive robust FO control structure (43) with an equivalently adjustable term
Lˆ (q̈, q˙ , q, p
˜,  ˆ ) ∈ R3 and an adaptive FO mechanism defined by
ˆ ,W

Lˆ (z ) = 
ˆ T g (z ) (46)

and
Dα 
ˆ = −βD2(α −1) g(z )eT M−1 (q ).
eq (47)

The flexible FO controller (43), combined with its adaptive FO estimator (46) and (47), well tracks the outputs q=[qa qu ]T
of TBC dynamics (6) and (7) to destination qd =[qad qud ]T in the sense of Mittag–Leffler stability even when the nonlinear
system (6) and (7) faces uncertain parameters p, unknown wind W and faults  of actuators.

Proof. In Theorem 1, the controller (24) requires the full knowledge of nonlinearities and faults of actuators. Although the
controller (24) is valid with parametric uncertainties and disturbances, it has no adaptation and no intelligence. Enhancing
the controller (24), the flexible controller (43) does not need the real information of the above-mentioned unfavourable
impacts. With the support of the adaptive mechanism (46) and (47) and without needing any physical measurement, the
system nonlinearities, winds, parametric uncertainties and faults will be identified and then provided to the controller (43).
Again, the main benefit here is to reduce the processing time because the system only identifies the lumped component
Lˆ (q̈, q˙ , q, p
˜,  ˆ ) instead of separately approximating many terms composed of Ba (q, q˙ ), Beq (q, q˙ ),Geq (q),Feq (q, q˙ ), p
ˆ ,W ˜ ,W
and . The adaptive mechanism (46) and (47) is constructed on the basis of neural estimation technique. Accordingly, we
utilise a neural network whose five inputs represented by a vector z = [ ea eu ]T ∈ R5 , one hidden layer with n nodes
and three outputs Lˆ (z ) ∈ R are approximation of unknown L(q̈, q˙ , q, p
3 ˜ , , W ). The input mapping converts inputs into a

814
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

scalar by using a matrix  ∈ Rn×3 of weights and is thus dependent on this weight matrix. After the inputs are mapped,
the signals are sent to activation functions to produce the outputs as approximations Lˆ (z ). The activation functions are
characterised by a vector g(z ) ∈ Rn of Gaussian functions as
  
z − b j 2
g(z ) = exp ( j = 1 − n) (48)
2c2j


with bj being medians, cj being standard deviations and z − b j 2 = 5i=1 (zi − b j ) indicating n Euclidean norms corre-
2

sponding to n vector (z − bj ).
The equivalently unknown component L(q̈, q˙ , q, p ˜ , , W ) of the controller (43) is estimated through mapping and acti-
vation, as shown in Eq. (46). Meanwhile,  is approximated using the relation (47). Here,  ˆ ∈ Rn×3 is an approximation of
 ∈ R , and β = diag(βi ) ∈ R
n ×3 n ×n is a matrix of adaptive parameters.
We consider a bounced positive Lyapunov candidate
1 T 1 1
V + Va + Vn = e e + eTu ρT ρeu + tr 
˜ T β−1 
˜ >0 (49)
2 2 2
Two first terms V and Va are chosen similar to (26) and (35) for verifying the convergence of global regime e and locally
unactuated error eu . The last term Vn is for checking the convergence of estimated value Lˆ . Differentiating Lyapunov (49) in
term FO α with respect to time, one obtains
1 α T 1 1
Dα (V + Va + Vn ) = D (e e ) + Dα eTu ρT ρeu + Dα tr 
˜ T β−1 
˜ (50)
2 2 2
The application of Lemmas 4 and 5 of [35] leads the Lyapunov derivative (50) to be

Dα (V + Va + Vn ) ≤ eT Dα e + (ρeu )T Dα (ρeu ) + tr 
˜ T β−1 Dα 
˜ (51)

Additionally, closed-loop dynamics is obtained when the dynamic model (15) suffers adaptive input (43)
 2 ( 1 −α )

ua (q ) L (z ) − Meq (q )D
Ua = J−1 ˆ 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne (52)

that is
 
ëa + D2(1−α ) 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne = M−1
eq (q ) L (z ) − L (q̈, q
ˆ ˜ , , W )
˙ , q, p (53)

or equivalently

D2α ea + 2λDα ea + λT λea + ρDα eu + λρ eu + ηsgne = M−1


eq (q )D
2(α −1 ) ˜
L (z ) (54)

which is reduced into

Dα e + λe + ρDα eu + λρ eu + ηsgne = M−1


eq (q )D
2(α −1 ) ˜
L (z ) (55)

where L˜ (z ) = Lˆ (z ) − L(q̈, q˙ , q, p
˜ , , W ) is the estimation error defined by mapping L˜ (z ) = 
˜ T g ( z ).
Combining closed-loop dynamics (55) with the Lyapunov inequality (51) yields

Dα (V + Va + Vn ) ≤ −eT λe − eT ηsgne + eT M−1


eq (q )D
2(α −1 ) ˜ ˜ T β−1 Dα 
L(z ) − eTu ρT λρ eu + tr  ˜ (56)

which can be rewritten as


 
Dα (V + Va + Vn ) ≤ −eT λe − eT ηsgne − eTu ρT λρ eu + tr eT M−1
eq (q )D
2(α −1 ) ˜ T ˜ T β−1 Dα 
 g (z ) +  ˆ (57)

The action of the adaptive FO estimator (47) leads expression (57) to be

Dα (V + Va + Vn ) ≤ −eT λe − eT ηsgne − eTu ρT λρ eu ≤ −λmin e2 − ηmin |e| − ρmin


λ
e u  (58)
λ = min(ρ 2 λ + ρ 2 λ , ρ 2 λ ). The application of Theorem 6.2 of [36] to expressions (49) and (58) shows that e, e
with ρmin 1 1 3 3 2 2 u
and  are bounced. That is, outputs q=[qa qu ]T approach qd =[qad qud ]T , and estimation 
˜ ˆ reaches  as time goes to infinity.
Because the control law (43) is structured using a fractional derivative, the adaptive mechanism (46) and (47) is also made
with a fractional derivative to yield a uniform control structure.

Remark 2. Guide for selecting control gains: The parameters of the controller (43) and the gains of the adaptive mechanism
(45) and (47) play individual roles. λ influences the speed of exponential convergence of actuated outputs, whilst ρ impacts
the stabilisation of unactuated outputs. Gain η affects the system output’s consistency: a small η can reduce the chattering
but increase the reaching time, while a large η improves the robustness with oscillation and instability trade-offs. Increasing
gain β speeds up the estimation.

815
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Table 1
Crane parameters.

Liebherr LTF 1060 crane Faults

mp = 300 kg; lt = 7 m; lb = 11 m; mt = 1200 kg;  = 10−5 diag(1.5, 1.2, 1.4 )


mb = 2200 kg; Jt = 13800 kgm2 ; Jb = 24000 kgm2 ; = 45 sin φ s + 15 cos φ˙ s + 20 cos φ˙ s˙ for Tf ≥ 30
1
kp = 28000 N/m; bp = 120 Ns/m; kl = 240000 N/m; = −0.8Fl for 25 ≤ Tf ≤ 30
2
bl = 450 Ns/m; ke = 27000 N/m; be = 180 Ns/m;
Rh = 0.4 m; Jh = 17090 kgm2 ; h0 = 3 m; a = 4; d = 5.5 m; 3 = 50 cos qs + 30 sin q˙ s + 15 cos q˙ s˙ for Tf ≥ 32

r = 5.4 m; n = 9 m; m = 0.3 m; k = 0.3 m; γ = 45o ; λ = 5o ; Winds


η = 0.97; g = 9.81 m/s2 ; q(0) = q0 ; q˙ (0 ) = 01×5 . qw = 250N/m2 ; A1 = 4.5 m2 ; k1 = 0.4; c1 = 0.73; n1 = 1;
χ 1 = 0.95; A2 = 2.5 m2 ; k2 = 0.3; c2 = 0.6; n2 = 1.1;
χ 2 = 0.8; A3 = 0.25 m2 ; k3 = 0.25; c3 = 0.3; n3 = 1.05;
χ 3 = 1.

Table 2
Parameters of controllers.

Robust adaptive FOC PI [10], PID [22] control

Controller: α = [0.9, 0.95, 1]; Kp = diag(1.2,1.4,1.7); Kd = diag(1.9,1.3,1.6);


λ = diag(0.9,0.35,0.6); ρ = [1 0; 0 1.1; 0 0]; Ki = = [0.8 0 0; 0 0.9 0;]T .
η = 10−3 diag(12,15,14). Extension of SMC [12]
Estimator: β = diag(2. 2,2.9,3.1,2.4,3.05); λ = diag(0.8,0.4,0.5); ρ = [1.1 0; 0 0.9; 0.3 0];
bj = [0.15 0.2 0.2 0.15 0.15]; K = diag(250,210,220).
cj = [−0.4 −0.2 0 0.4 0.25]; (0) = 06 × 3 .

4. Application to a practical TBC

4.1. Simulation

We apply the control system composed of a robust controller (43) and its neural estimator (46) and (47) to a practical
crane, namely, Liebherr LTF 1060, through simulation. To show the effectiveness of the proposed control system, the com-
parison with other approaches [12,22] is discussed. The PID controller of previous article [22] is redesigned as follows:
 
t
UPID
a ua (q ) Ba (q, q
= J−1 ˙ )q˙ a + Beq (q, q˙ )q˙ u + Geq (q ) + Feq (q, q˙ ) − Meq (q )(K p ea + Kd e˙ a + Ki eu dt ) (59)
0

K 0 Ki3 T
where K p = diag(K p1 , K p2 , K p3 ), Kd = diag(Kd1 , Kd2 , Kd3 ), and Ki = [ i1 ] .
0 Ki2 0
The SMC by Vázquez et al. [12] is applied only for telescoping the boom. We enhance this algorithm for five motions
q ∈ R5 as follows:
Ua = Ueq + Ksgns (60)
where s is a sliding surface defined by
s = e˙ a + λea + ρeu (61)
Ueq is acquired from the condition s˙ + λs = 0, and K=diag(K1 ,K2 ,K3 ) refers to switching gains. The parameters necessary
for the simulation are indicated in Tables 1 and 2.
The influences of winds and faults on the control system are considered. Their characteristics are presented in Table 1.
Given that winds are lumped forces acting on the mass centres of the base section, top section and payload, [wi ]T =[wB wC
wp ]T . Each component is defined by
wi = Akcnχ qw (62)
where A is the area of wind resistance, k is a coefficient reflecting the variation in wind pressure in accordance with the
height, c is an aerodynamic coefficient, n is an overload coefficient, χ reflects the structural tightness, and qw is the wind
intensity. Winds wi act on the crane components from the 20 to 25th seconds.
Faults  of the three actuators imply that 1 has occurred in the hoisting motor since the 30th second, 2 reduces the
efficiency of the boom-luffing cylinder by 80% from the 25 to 30th seconds, and 3 has appeared at the telescopic cylinder
since the 32nd second.
From the initial position q0 =[φ 0 ψ 0 q0 − q s0 + s θ 0 ]T =[0 π /6 1− q 5+ s 0]T , the controllers track the crane outputs
to destination qd =[2a/Rh π /3 1.5 3 0]T . Here, q and s are the static deformations of the telescopic cylinder and load-
carrying rope that are, respectively given by
m pg
s= (63)
ak p

816
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 2. Rotation of hoisting drum.

Fig. 3. Luffing angle of boom.

Fig. 4. Expansion of telescopic boom.

and
1
 mp
 
q= (m p + mt ) sin ψ0 + g − Wc cos ψ0 (64)
ke aη

4.2. Results and discussion

Figures. 2–6 show the crane responses tracked by neural FOC. The specifications of responses comprising maximum over-
shoots Mp , settling time ts , steady-state error ess , maximum deviation due to wind Aw and maximum deviation due to fault
Af measured in Figs. 3–6 are represented in Table 3. The controller well tracks the hoisting drum, boom rotation and boom
expansion to destination concurrently whilst lifting the payload to the desired rope length, maintaining the cargo swing
small (θ min ≈−0.7°, θ max ≈1.4°) and suppressing it at the load destination. Various FOs α =[0.9 0.95 1] create the derivations
of transient responses; however, steady responses converge at unique references. Increasing α reduces the settling time and

817
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 5. Lifting the payload.

Fig. 6. Payload swing.

Table 3
Characteristics of crane responses under the neural FOC controller.

FOC Controller Boom luff (Fig. 3) Boom extension (Fig. 4) Load hoist (Fig. 5) Load swing (Fig. 6)

FO = 1 Mp 0.00 0.00 0.00 1.56


ts 12 15 13 14
ess 0 0 0 5.10−3
Aw 0.240 0.007 0.036 0.144
Af 0.090 0.006 0.055 0.030
FO = 0.95 Mp 0.28 0.00 0.00 1.44
ts 13 16 14 16
ess 0 0 0 7.10−3
Aw 0.150 0.008 0.035 0.139
Af 0.170 0.007 0.058 0.06
FO = 0.9 Mp 0.29 0.00 0.00 1.34
ts 14 19 18 17
ess 0 0 0 8.10−3
Aw 0.030 0.009 0.031 0.110
Af 0.330 0.008 0.061 0.135

transient phase, but it is insignificant and may cause overshoot. Given that winds act on the boom and payload in accor-
dance with the horizontal direction, their effects on the longitudinal motion of payload (Fig. 5) and rolling cable (Fig. 2)
are insignificant. The neural adaptation mechanism is combined with the controller to approximate and compensate for
winds and faults of actuators. As a result, the influence of faults and winds on the boom luff and load hoist is considerably
suppressed, as demonstrated in Figs. 3 and 5. Winds and faults only cause a minimal deviation to the boom luff since the
20th second (Fig. 3). Despite the usefulness of the adaptive mechanism, the impact of winds and faults is still observed in
boom retraction (Fig. 4) and cargo swing (Fig. 6). Winds cause deviation from the 20 to 25th seconds, whilst faults induce
deformation from the 25 to 35th seconds. The dynamic reaction between the hoisting drum and payload motion through
the elasticity of the handling rope and the flexibility of the telescopic cylinder contributes to increased impacts of winds
and faults. Figs. 7–9 illustrate the control inputs of actuators. Reducing FOs may increase the chattering of control inputs.
Depicted in Fig. 10 are estimations of equivalence Lˆ (z ). It is not important for estimation Lˆ to converge to the truth value L.

818
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 7. Torque of hoisting drum.

Fig. 8. Pushing force of luff cylinder.

Fig. 9. Force of telescopic cylinder.

The main objective of the control system is to track and stabilise the outputs. The controller is combined with the estimator
to realise this goal. Thus, Lˆ can approach various destinations as long as the system is stabilised.

4.3. System sensitivity

A TBC holds many uncertain parameters, such as load mass mp and frictions bp , bl , be . We investigate the sensitivity of
crane responses tracked by neural FOC for case FO=0.95 when a crane faces two cases of parametric variations: mp =[100
−50]%, bp =[−20 20]%, bl =[10 −20]% and be =[−10 5]%. The outputs depicted in Fig. 11 remain consistent despite the
large variations in four crane parameters. Luffing the boom and hoisting the load seem insensitive to varying parameters.
Boom extension and load swing have a small curve variation at transient states but are generally concise at steady states.

819
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 10. Approximations of Lˆ (q̈, q˙ , q, p


˜,  ˆ ).
ˆ ,W

Fig. 11. Robustness of outputs.

820
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Fig. 12. Crane responses using extended PID control [22].

Fig. 13. Crane responses using extended SMC [12].

4.4. Comparison

Figures. 12 and 13 depict the crane outputs under the action of PID [22] and SMC [12] control. The specifications of
these outputs measured in Figs. 12 and 13 are represented in Table 4. The simulation of PI and PID controllers [12,22]
(59) is unsuccessful when winds and faults are introduced into the dynamic model. In fact, PI and PID are not useful in the
presence of parametric variation, faults and winds. Although their control structures are simple, they are very sensitive to

821
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Table 4
Characteristics of crane responses under the SMC [12] and PID [22] controllers.

Boom
Controllers Boom luff extension Load hoist Load swing

Extended Mp 0.0 0.0 0.0 1.7


SMC ts 15 13 14 18
(Fig. 12) ess 0 0 0 0
Aw 0.260 0.095 0.000 0.110
Af 0.90 0.10 0.00 0.13
PID Mp 6.28 0.01 0.00 19.73
(Fig. 13) ts 12 10 11 19
ess 0.05 0.06 0.00 0.06
Aw X X X X
Af X X X X

(X) means no validation for winds and faults.

uncertainties and have no robustness. A small variation in impacts may lead to the crane’s instability. The crane responses
tracked by the PID controller without the action of varying parameters, winds and faults are depicted in Fig. 12. The large
boundary of payload swing, θ min ≈−20° and θ max ≈19°, remains, and the many oscillations before vanish. Overshoots are
observed in boom swing and extension. The industrial application of PID controllers has been verified in articles [10,22].
As shown in Fig. 13, the extended SMC seems better than PID, in which the payload swing is kept very small, θ min =−0.2°
and θ max =0.8°, and no overshoots exist in responses. In fact, the extension (60) and (61) of SMC [12] is effective despite
uncertainties, and the controller keeps the consistency of outputs despite winds and faults. Relying on the switched com-
ponent of the controller simply maintains the system’s robustness without intelligence and adaptation. The control frame is
stiffly fixed with no flexible variation to adapt to uncertain job sites.
Whilst the above-mentioned approaches are based on feedback principle, time-delay and command shaping controls
[21] are open-loop approaches. Without the need for any sensor for measuring feedback signals, the controller still works
well; however, it is only valid for small variations in crane parameters. It is rather sensitive to winds and faults. The strong
winds and significant faults may make the crane unstable if this approach is used.
In principle, our approach is superior to the others. It holds both robust and adaptive features and tracks the crane
outputs whilst concurrently estimating and suppressing the uncertainties and influences. The crane responses (Figs. 2–6) of
our approach are not better than the others though [12,22]. The reason is that it conducts many duties at the same time
and suffers many influences. Meanwhile, PI [10] and PID [22] controllers do not consider the impact of winds and faults. PI
[10], PID [22], SMC [12] and input shaping control [21] only focus on tracking duty without identification and estimation
of uncertainties and disturbances. The others [10,12,21,22] utilise the fixed first order of derivative and integral. On the
contrary, our controller uses variable FOs, resulting in a flexible control structure.

Remark 3. Note to Practitioners: In the next step, the control structure should be tested on a laboratory-scale crane, and
the control algorithm should be run on a Xilinx® microprocessor. Then, this control structure could be implemented on
an industrial telescope crane, and the control gains could be tuned to optimise the crane performance. Lastly, the control
algorithm could be integrated into the PLC and CAN versions of industrial controllers.

5. Conclusion

We proposed a new control system that shows high applicability to industrial TBCs. The effectiveness of controller was
validated by simulation for three cases of fractional-orders 0.9, 0.95, and 1. Boom luffing, boom expansion and payload
hoist are tracked to destinations precisely, whilst payload swing is kept small during the transport and vanishes at the
load destination. The oscillations due to elasticity (payload oscillation and boom swing) are almost suppressed. The control
system achieves both robust and adaptive features. Only one adaptive mechanism is used, but it detects and estimates
structural uncertainties, parametric variations, winds and actuator faults concurrently. The control system combines the
advantages of FOC, an SMC-like method and a neural network-based adaptation. However, the controller holds the complex
structure due to simultaneously conducting many objectives.
In the future work, we will enhance the dynamic model by supplementing the dynamic features of electric motors,
hydraulic motors and cylinders. We will integrate deep and reinforcement learning into control algorithms to make the
crane more intelligent and mimic human behaviour. Experimental validation will be included.

Data availability statement

The data sets generated during the current study are not publicly available due to the request of the sponsor but are
available from the corresponding author upon reasonable request.

822
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

Acknowledgments

The Authors are sincerely grateful to Professor Lev V. Idels, the handling Editor, and the Reviewers for their valuable
suggestions, which greatly improved the quality of the paper. This work is supported by Vietnam Maritime University.

Appendix

The entries of inertial matrixes include


2 2
m11 = Jh , m22 = Jt + Jb + 0.25mb lb2 + m p [0.5(lt + lb ) + q] + m p k2 + mt (0.5lb + q) , m33 = m p + mt ,

m44 = m p , m55 = m p s2 , m23 = m32 = m p k, m24 = m42 = m p {k sin (θ − ψ ) − [0.5(lt + lb ) + q] cos (θ − ψ )},

m25 = m52 = m p s{k cos (θ − ψ ) + [0.5(lt + lb ) + q] sin (θ − ψ )},

m34 = m43 = m p sin (θ − ψ ), m35 = m53 = m p s cos (θ − ψ ),

m12 = m13 = m14 = m15 = m21 = m31 = m41 = m45 = m51 = m54 = 0.

The components of two damping vectors compose of


 
b1 = b p R2h , b2 = [m p lt + (m p + mt )lb + (2mt + m p )q]q˙ + bl n2 d2 sin (γ − λ + ψ )/ll2 (ψ ) ,
2

b3 = b p + be , b4 = b p a2 , b5 = 2m p ss˙ ,

The elements of two gravitational vectors are given by

Rh m p
g1 = g, g2 = {m p [0.5(lt + lb ) + q] + 0.5mb lb + mt (0.5lb + q)}g cos ψ ,
a

m pg
g3 = (m p + mt )g sin ψ + , g4 = −m p g cos θ + m p g, g5 = m p gs sin θ .
a

The nonlinearities include

f1 = b p Rh as˙ + b p Rh q˙ + k p Rh [Rh (φ − φ0 ) + a(s − s0 ) + (q − q0 )],

2s˙ θ˙ [0.5(l + l ) + q] − ksθ˙ θ˙ − ψ˙  sin (θ − ψ ) 


 t b

f2 = m p + sθ˙ 2 [0.5(lt + lb ) + q] + ks˙ 2θ˙ − ψ˙ cos (θ − ψ )
 
−k sψ˙ θ˙ sin (θ − ψ ) − ψ˙ s˙ cos (θ − ψ ) − g sin ψ
+kl [ lls + ll (ψ ) − ll (ψ0 )]nd sin(γ − λ + ψ )/ll (ψ ),

f3 = ke ( les + q − q0 ) + k p [Rh (φ − φ0 ) + a(s − s0 ) + (q − q0 )] + b p Rh φ˙ + b p as˙ 


−{0.5m p lt + (m p + mt )(0.5lb + q)}ψ˙ 2 + m p 2s˙ θ˙ cos (θ − ψ ) − sθ˙ 2 sin (θ − ψ ) ,

f4 = b p a Rh φ˙ + q˙ + k p a[Rh (φ − φ0 ) + (q − q0 )] − m p kψ˙ 2 + 2q˙ ψ˙ cos (θ − ψ )


+k p a2 (s − s0 ) − m p ψ˙ 2 [0.5(lt + lb ) + q] sin (θ − ψ ) − m p sθ˙ 2 ,

 
f5 = m p sψ˙ 2q˙ + kψ˙ sin (θ − ψ ) − ψ˙ [0.5(lt + lb ) + q] cos (θ − ψ ) .

823
L.A. Tuan and L.V. Duong Applied Mathematical Modelling 108 (2022) 807–824

References

[1] Y. Liu, H. Yu, A survey of underactuated mechanical systems, IET Control Theory Appl. 7 (7) (2013) 921–935, doi:10.1049/iet-cta.2012.0505.
[2] B. Alexey, K. Alexander, K. Yury, K. Anton, Stabilization of the motion of a spherical robot using feedbacks, Appl. Math. Model. 69 (2019) 583–592,
doi:10.1016/j.apm.2019.01.008.
[3] G. Qin, et al., Adaptive trajectory control of an under-actuated snake robot, Appl. Math. Model. 106 (2022) 756–769, doi:10.1016/j.apm.2022.02.001.
[4] A. Zhang, X. Lai, M. Wu, J. She, Global stabilization of underactuated spring-coupled three-link horizontal manipulator using position measurements
only, Appl. Math. Model. 39 (7) (2015) 1917–1928, doi:10.1016/j.apm.2014.10.010.
[5] A.T. Le, S.G. Lee, 3D cooperative control of tower cranes using robust adaptive techniques, J. Frankl. Inst. Eng. Appl. Math. 354 (18) (2017) 8333–8357,
doi:10.1016/j.jfranklin.2017.10.026.
[6] L.A. Tuan, Fractional-order fast terminal back-stepping sliding mode control of crawler cranes, Mech. Mach. Theory 137 (2019) 297–314, doi:10.1016/j.
mechmachtheory.2019.03.027.
[7] R. Mijailović, Modelling the dynamic behaviour of the truck-crane, Transport 26 (4) (2011) 410–417, doi:10.3846/16484142.2011.642946.
[8] I.A. Wójcik, Ł. Drag,˛ M. Metelski, K. Nadratowski, S. Wojciech, A 3D model for static and dynamic analysis of an offshore knuckle boom crane, Appl.
Math. Model. 66 (2019) 256–274, doi:10.1016/j.apm.2018.09.006.
[9] I.A. Martin, R.A. Irani, Dynamic modeling and self-tuning anti-sway control of a seven degree of freedom shipboard knuckle boom crane, Mech. Syst.
Signal Proc. 153 (2021) 107441, doi:10.1016/j.ymssp.2020.107441.
[10] A. Urbaś, J. Kłosiński, K. Augustynek, The influence of the PID controller settings on the motion of a truck-mounted crane with a flexible boom and
friction in joints, Control Eng. Pract. 103 (2020) 104610, doi:10.1016/j.conengprac.2020.104610.
[11] K. Watanabe, M. Yoshikawa, J. Ishikawa, Damping control of suspended load for truck cranes in consideration of second bending mode oscillation, in:
Proceedings of the 44th Annual Conference of the IEEE Industrial Electronics Society, 2018, pp. 4561–4568, doi:10.1109/iecon.2018.8591232.
[12] C. Vázquez, S. Aranovskiy, L. Freidovich, L. Fridman, Second order sliding mode control of a mobile hydraulic crane, in: Proceedings of the 53rd IEEE
Conference on Decision and Control, 2014, pp. 5530–5535, doi:10.1109/cdc.2014.7040254.
[13] Y. Qian, Y. Fang, B. Lu, Adaptive repetitive learning control for an offshore boom crane, Automatica 82 (2017) 21–28, doi:10.1016/j.automatica.2017.04.
003.
[14] J. Neupert, E. Arnold, K. Schneider, O. Sawodny, Tracking and anti-sway control for boom cranes, Control Eng. Pract. 18 (1) (2010) 31–44, doi:10.1016/
j.conengprac.20 09.08.0 03.
[15] A. Trabka,
˛ Dynamics of telescopic cranes with flexible structural components, Int. J. Mech. Sci. 88 (2014) 162–174, doi:10.1016/j.ijmecsci.2014.07.009.
[16] B. Posiadała, B. Skalmierski, L. Tomski, Motion of the lifted load brought by a kinematic forcing of the crane telescopic boom, Mech. Mach. Theory 25
(5) (1990) 547–556, doi:10.1016/0 094-114X(90)90 068-U.
[17] B. Posiadała, Influence of crane support system on motion of the lifted load, Mech. Mach. Theory 32 (1) (1997) 9–20, doi:10.1016/0094-114X(96)
0 0 044-4.
[18] B. Posiadala, D. Cekus, Discrete model of vibration of truck crane telescopic boom with consideration of the hydraulic cylinder of crane radius change
in the rotary plane, Autom. Constr. 17 (3) (2008) 245–250, doi:10.1016/j.autcon.2007.05.004.
[19] D. Cekus, R. Gnatowska, P. Kwiatoń, Impact of wind on the movement of the load carried by rotary crane, Appl. Sci. 9 (18) (2019) 3842, doi:10.3390/
app9183842.
[20] D. Cekus, P. Kwiatoń, Effect of the rope system deformation on the working cycle of the mobile crane during interaction of wind pressure, Mech.
Mach. Theory 153 (2020) 104011, doi:10.1016/j.mechmachtheory.2020.104011.
[21] J.Y. Park, P.H. Chang, Vibration control of a telescopic handler using time delay control and commandless input shaping technique, Control Eng. Pract.
12 (6) (2004) 769–780, doi:10.1016/j.conengprac.2003.09.005.
[22] J. Činkelj, R. Kamnik, P. Čepon, M. Mihelj, M. Munih, Closed-loop control of hydraulic telescopic handler, Autom. Constr. 19 (7) (2010) 954–963,
doi:10.1016/j.autcon.2010.07.012.
[23] H. Fujita, H. Sugiyama, Development of flexible telescopic boom model using absolute nodal coordinate formulation sliding joint constraints with
LuGre friction, Theor. Appl. Mech. Lett. 2 (6) (2012) 063005, doi:10.1063/2.1206305.
[24] H.C. Pedersen, T.O. Andersen, B.K. Nielsen, Comparison of methods for modeling a hydraulic loader crane with flexible translational links, J. Dyn. Syst.
Meas. Control Trans. ASME 137 (10) (2015) 101012, doi:10.1115/1.4030801.
[25] W. Sochacki, The dynamic stability of a laboratory model of a truck crane, Thin-Walled Struct 45 (10–11) (2007) 927–930, doi:10.1016/j.tws.2007.08.
023.
[26] M. Savković, M. Gašić, G. Pavlović, R. Bulatović, N. Zdravković, Stress analysis in contact zone between the segments of telescopic booms of hydraulic
truck cranes, Thin Walled Struct. 85 (2014) 332–340, doi:10.1016/j.tws.2014.09.009.
[27] D. Shang, X. Li, M. Yin, F. Li, Dynamic modeling and fuzzy compensation sliding mode control for flexible manipulator servo system, Appl. Math.
Model. (2022) In Press, doi:10.1016/j.apm.2022.02.035.
[28] H. Richter, D. Simon, W.A. Smith, S. Samorezov, Dynamic modeling, parameter estimation and control of a leg prosthesis test robot, Appl. Math. Model.
39 (2) (2015) 559–573, doi:10.1016/j.apm.2014.06.006.
[29] S.B. Chen, et al., The effect of market confidence on a financial system from the perspective of fractional calculus: Numerical investigation and circuit
realization, Chaos Solitons Fractals 140 (2020) 110223, doi:10.1016/j.chaos.2020.110223.
[30] H. Jahanshahi, et al., Simulation and experimental validation of a non-equilibrium chaotic system, Chaos Solitons Fractals 143 (2021) 110539, doi:10.
1016/j.chaos.2020.110539.
[31] S.S. Zhou, et al., Discrete-time macroeconomic system: Bifurcation analysis and synchronization using fuzzy-based activation feedback control, Chaos
Solitons Fractals 142 (2021) 110378, doi:10.1016/j.chaos.2020.110378.
[32] S.-B. Chen, et al., Optimal control of time-delay fractional equations via a joint application of radial basis functions and collocation method, Entropy
22 (11) (2020) 1213, doi:10.3390/e22111213.
[33] C.J. Zuñiga-Aguilar, et al., Robust control for fractional variable-order chaotic systems with non-singular kernel, Eur. Phys. J. Plus 133 (13) (2018),
doi:10.1140/epjp/i2018-11853-y.
[34] J.F. Li, et al., On the variable-order fractional memristor oscillator: data security applications and synchronization using a type-2 fuzzy disturbance
observer-based robust control, Chaos Solitons Fractals 145 (2021) 110681, doi:10.1016/j.chaos.2021.110681.
[35] M.A.D. Mermoud, N.A. Camacho, J.A. Gallegos, R.C. Linares, Using general quadratic lyapunov functions to prove lyapunov uniform stability for frac-
tional order systems, Commun. Nonlinear Sci. Numer. Simul. 22 (1) (2015) 650–659, doi:10.1016/j.cnsns.2014.10.008.
[36] Y. Li, Y.Q. Chen, I. Podlubny, Stability of fractional-order nonlinear dynamic systems: lyapunov direct method and generalized Mittag-Leffler stability,
Comput. Math. Appl. 59 (5) (2010) 1810–1821, doi:10.1016/j.camwa.2009.08.019.

824

You might also like