Vous êtes sur la page 1sur 10

Control Engineering Practice 86 (2019) 1–10

Contents lists available at ScienceDirect

Control Engineering Practice


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

Predictive functional control of an axis positioning system with an


estimator-based internal model
Toshiyuki Satoh a ,∗, Naoki Saito a , Jun-ya Nagase b , Norihiko Saga c
a Department of Intelligent Mechatronics, Akita Prefectural University, Yurihonjo, Akita, Japan
b
Department of Mechanical and Systems Engineering, Ryukoku University, Otsu, Shiga, Japan
c
Department of Human System Interaction, Kwansei Gakuin University, Sanda, Hyogo, Japan

ARTICLE INFO ABSTRACT


Keywords: This paper deals with an application of the predictive functional control with a state estimator-based internal
Predictive functional control model (PFC_ EBIM). The PFC_ EBIM has been shown to be effective in simulation. However, neither detailed
Model-based control experimental validation nor comparison with other controllers has been reported thus far. Here, the PFC_ EBIM
Axis positioning system
is implemented in a single-axis positioning system, and a few experimental tests are conducted. Tracking
Mechatronics
performance of the PFC_ EBIM, standard PFC, and P – PI control for both smooth and non-smooth reference
Motion control
Real-time control
signals are compared. The experimental results prove the effectiveness of the PFC_ EBIM.

1. Introduction heat exchanger (Abdelghani-Idrissi, Arbaoui, Estel, & Richalet, 2001;


Richalet, Darure, & Mallet, 2014), a tandem cold rolling mill (Friebel,
The model predictive control (MPC), which emerged from the pro- Zabet, Haber, & Jelali, 2015), a conductivity process (Zabet & Haber,
cess control area, is a practical model-based control strategy (Camacho 2017), a boost converter (Anzehaee, Behnam, & Hajihosseini, 2018),
& Bordons, 2007; Maciejowski, 2002; Qin & Badgwell, 2003; Richalet, an autonomous underwater vehicles (Muñoz, Sellier, & Castro, 2018),
1993) and has recently become popular in many control fields including and a batch reactor (Bouhenchir, Cabassud, & Lann, 2006; Dovžan
robotics and mechatronics. The advantages of the MPC include its & Škrjanc, 2010). Apart from such applications, contributions to the
ability to handle multiple-input–multiple-output systems, impose con- theory of the PFC are also in progress (Rossiter, 2002, 2004; Rossiter,
straints on states and control input, and execute the optimal operation Haber, & Zabet, 2016; Zhang, Xue, Wang, & Zhang, 2012; Zhang, Yang,
with the aid of numerical optimization algorithms.
Zong, Wu, & Zhang, 2011).
Predictive functional control (PFC) (Kuntze, Jacubasch, Richalet, &
In contrast to other MPC strategies, the PFC predicts the future plant
Arber, 1986; Richalet, 1993; Richalet et al., 1987; Richalet & O’Donovan,
output only on the basis of an internal independent model; no state
2009) is a simple model predictive control strategy that is mainly
estimator is usually required. Hence the states of the internal model
intended for single-input–single-output systems. The computational
could be different from those of the actual plant in the presence of
load of the PFC is light as compared to the standard MPC, since no
online optimization is required. Hence, unlike the standard MPC, the disturbances and/or uncertainties. Nevertheless, thanks to the integral
PFC has been used to control mechatronic systems from the initial stage action in the PFC, the steady-state error to a step-wise reference be-
of its development (Kuntze et al., 1986; Richalet et al., 1987). Many comes zero, as long as the closed-loop stability is ensured. However,
successful PFC-based applications have been reported since then. It in the case of integrating plants (i.e., plants with a pole at 𝑧 = 1),
has been used, for example, in the control of a magnetic suspension an offset arises to step-wise disturbance applied to the control input.
system (Lepetič, Škrjanc, Chiacchiarini, & Matko, 2003), the H4 parallel This is because the PFC controller has an integrator (i.e., pole at
robot (Vivas & Poignet, 2005), a brushless DC servo motor (Dieulot, 𝑧 = 1) and a zero at 𝑧 = 1 at the same time to cancel the plant
Benhammi, Colas, & Barre, 2008), an active queue management in pole at 𝑧 = 1. As a result, pole-zero cancellation occurs between the
TCP/IP networks (Bigdeli & Haeri, 2009), a pneumatic artificial mus- numerator and denominator polynomials, and the PFC controller loses
cle (Satoh, Saito, & Saga, 2010), a PMSM servo system (Liu & Li, an integral action. This is often the case with optimal control-based
2012), a table drive system (Satoh, Kaneko, & Saito, 2012), a tendon- methods. To overcome this drawback, the authors developed a novel
driven balloon actuator (Nagase, Hamada, Satoh, Saga and Suzumori, PFC scheme called the predictive functional control with an estimator-
2013; Nagase, Satoh, Saga and Suzumori, 2013), a counter-current based internal model (PFC_ EBIM) (Satoh, Saito, Nagase, & Saga, 2016),

∗ Corresponding author.
E-mail address: tsatoh@akita-pu.ac.jp (T. Satoh).

https://doi.org/10.1016/j.conengprac.2019.02.006
Received 24 August 2018; Received in revised form 14 December 2018; Accepted 19 February 2019
Available online xxxx
0967-0661/© 2019 Elsevier Ltd. All rights reserved.
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

using an active disturbance estimation based on an augmented system


and a state estimator. In the design of the PFC_ EBIM, a prediction
estimator is utilized not merely to estimate current states, but also to
compute the future free and forced responses subject to disturbance
necessary to derive the optimal control law. Hence no additive mech-
anism such as the disturbance observer is required, and disturbance
rejection is accomplished indirectly. Several numerical simulation tests
were conducted to evaluate the effectiveness of the PFC_ EBIM (Satoh
et al., 2016). Also, a simple experimental application of the PFC_ EBIM
was presented in Satoh, Saito, Nagase, and Saga (2017). However,
neither detailed experimental validation nor comparison with other
controllers such as the standard PFC or the P – PI controller has been
reported in the literature.
This paper aims to implement the PFC_ EBIM on a single-axis posi-
tioning system and evaluate its effectiveness experimentally. Single-axis
positioning systems, also known as single-axis table drive systems, are
widely utilized in industries because of their simplicity and applica- Fig. 1. Concept of predictive functional control.
bility. They are used individually or in combination with two or more
systems to generate two- or three-dimensional movement. The position-
ing accuracy required for the positioning systems ranges from the order PFC is computed to minimize the error between the reference trajectory
of centimeters to micrometers. A simple control strategy for single- 𝑦𝑅 and the predicted plant output 𝑦̂𝑃 at the given coincidence points.
axis positioning systems is the P – PI control, where the PI controller Hence, in the PFC, the following performance index is usually used:
and P controller are located in the velocity and position-feedback loop,
𝑛ℎ
respectively (Matsubara, 2008). ∑ { ( ) ( )}2
𝐽 (𝑘) ∶= 𝑦̂𝑃 𝑘 + ℎ𝑗 − 𝑦𝑅 𝑘 + ℎ𝑗 . (2)
The basic procedure of implementation can be summarized as fol- 𝑗=1
lows. First, an augmented state-space model including disturbance is ( )
constructed by using a simplified linear model of the plant composed Therein, ℎ𝑗 ∈ N 𝑗 = 1, 2, … , 𝑛ℎ are the time at which coincidence
of a first-order lag system plus an integrator. The free and forced points are defined, and 𝑛ℎ ∈ N is the number of coincidence points,
responses subject to disturbance are then computed and, by summing where N is the set of all natural numbers. The predicted output 𝑦̂𝑃 is
them, the predicted output is computed. Finally, the optimal control computed as,
law can be analytically derived from the predicted output under a ( ) ( ) ( )
𝑦̂𝑃 𝑘 + ℎ𝑗 = 𝑦̂𝑈 𝐹 𝑘 + ℎ𝑗 + 𝑦̂𝐹 𝑘 + ℎ𝑗 , (3)
quadratic performance index. It is assumed that there are more than
two coincident points on the reference trajectory, and that the control where 𝑦̂𝑈 𝐹 ∈ R is the free (unforced) output response, and 𝑦̂𝐹 ∈ R
input is composed of the sum of weighted basis functions, which is a is the forced output response. In the PFC scheme, the control input is
common custom in the PFC. structured as a linear combination of basis functions as follows:
The organization of the paper is as follows. Section 2 reviews 𝑛𝐵

the PFC algorithm using a state estimator-based internal model. In 𝑢 (𝑘 + 𝑖) ∶= 𝜇𝑙 (𝑘) 𝑈𝐵𝑙 (𝑖) , (4)
Section 3, the experimental apparatus and its modeling are described. 𝑙=1
Section 4 introduces the compared position controllers used in the ( )
where 𝑈𝐵𝑙 ∈ R 𝑙 = 1, 2, … , 𝑛𝐵 are the predefined basis functions,
experiment. In Section 5, the experimental results are presented, and ( )
and 𝜇𝑙 ∈ R 𝑙 = 1, 2, … , 𝑛𝐵 are unknown weights. Then the optimal
the effectiveness of the PFC_ EBIM is validated. Section 6 contains our
control input trajectory that minimize the performance index 𝐽 (𝑘) can
conclusions.
be obtained by setting 𝜕𝐽 (𝑘) ∕𝜕𝝁 (𝑘) = 0 where
( )𝑇
2. Predictive functional control with a state estimator-based in- 𝝁 ∶= 𝜇1 (𝑘) 𝜇2 (𝑘) ⋯ 𝜇𝑛𝐵 (𝑘) . (5)
ternal model
Once we have obtained the input trajectory, we will apply its first
This section presents a summary of the standard PFC (Kuntze et al., element to the plant. At the next sampling time, we repeat the above-
1986; Richalet, 1993; Richalet et al., 1987; Richalet & O’Donovan, mentioned procedure. For details, consult Appendix A in Abu el Ata-
2009), the state estimator, and the PFC with a state estimator-based Doss, Fiani, and Richalet (1991).
internal model (PFC_ EBIM) (Satoh et al., 2016).
2.2. State estimator
2.1. Predictive functional control

A state estimator (sometimes called a state observer) is a mecha-


The concept of the standard PFC is summarized in Fig. 1, which
nism which computes an estimate of the entire state vector using the
shows a given set-point 𝑐 (𝑘) ∈ R at time 𝑘, a predicted plant output
measurements of a system (Franklin, Powell, & Workman, 1998). There
𝑦̂𝑃 ∈ R, a reference trajectory 𝑦𝑅 , and coincidence points. For simplic-
ity, the set-point is assumed to be constant. The choice of the number of are two types of full-order state estimators: a prediction estimator and
coincidence points is arbitrary, and, as an example, three coincidence a current estimator. A brief overview of a prediction estimator is given
points are depicted in Fig. 1. The reference trajectory, which indicates in the following since it is utilized in the PFC_ EBIM.
the desired future output trajectory towards the set-point, is defined as, Suppose that a single-input–single-output discrete-time state-space
representation of a plant is given follows:
( ) {
𝑦𝑅 (𝑘 + 𝑖) ∶= 𝑐 (𝑘 + 𝑖) − 𝛼 𝑖 𝑐 (𝑘) − 𝑦𝑃 (𝑘) , 𝑖 = 0, 1, … , (1) 𝒙 (𝑘 + 1) = Φ𝒙 (𝑘) + 𝜸𝑢 (𝑘) ,
(6)
𝑦 (𝑘) = 𝒉𝒙 (𝑘) ,
where 𝛼 ∈ R is a parameter that is chosen as 0 < 𝛼 < 1, and 𝑦𝑃 (𝑘) is the
plant output at time 𝑘. At each sampling time, the reference trajectory is where 𝒙 ∈ R𝑛 is the state vector, 𝑢 ∈ R is the control input, and
re-initialized using the measured plant output. The control input of the 𝑦 ∈ R is the plant output. We assume that the system given in Eq. (6)

2
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

is observable. Then a prediction estimator is given by the following 2.4. Control law of PFC_ EBIM
equation (Franklin et al., 1998):
The free and forced responses must be considered to derive the
𝒙̂ (𝑘 + 1) = Φ𝒙̂ (𝑘) + 𝜸𝑢 (𝑘) + 𝒍𝑝 [𝑦 (𝑘) − 𝒉𝒙̂ (𝑘)] , (7) optimal control law in the PFC scheme. When the current sampling
instant is 𝑘, the predicted 𝑖-step-ahead free response is written as,
where 𝒙̂ ∈ R𝑛 is the estimate of 𝒙, and 𝒍𝑝 ∈ R𝑛 is the estimator gain
vector. As the name implies, a measurement at time 𝑘 results in an 𝑦̂𝑈 𝐹 (𝑘 + 𝑖) = 𝒄 𝒂 𝑨𝑖𝒂 𝒙̂ 𝒂 (𝑘) . (16)
estimate of the state vector that is valid at time 𝑘 + 1. Now we define
If the predictive estimator given in Eq. (15) is utilized as the internal
the error in the estimate as,
model, then the total forced response is obtained as the result of the
𝒆 ∶= 𝒙 − 𝒙.
̂ (8) superposition of the one originated from 𝑢 over that originated from 𝜖.
The control input is assumed to be composed of the sum of weighted
Then a difference equation describing the behavior of 𝒆 is given by, basis functions in the PFC scheme as given in Eq. (4). The following
[ ] time-dependent polynomial basis is utilized in this study:
𝒆 (𝑘 + 1) = Φ − 𝒍𝑝 𝒉 𝒆 (𝑘) . (9)
𝑈𝐵𝑙 (𝑖) = 𝑖𝑙−1 . (17)
It can be understood from Eq. (9) that the error 𝒆 will converge to
The polynomial basis has been used in the design of the standard PFC.
zero for any value of 𝒆 (0) if Φ − 𝒍𝑝 𝒉 is a stable matrix. In general, the
Then the forced response originated from 𝑢 is computed as,
estimator gain vector 𝒍𝑝 is chosen such that the poles of the system in
𝑛𝐵
Eq. (9) are sufficiently fast. ∑
𝑦̂𝐹1 (𝑘 + 𝑖) = 𝜇𝑙 (𝑘) 𝑦1𝐵𝑙 (𝑖) , (18)
𝑙=1

2.3. Augmented internal model for PFC_ EBIM where 𝑦1𝐵𝑙 (𝑖) = 𝑖−1 𝑖−1−𝑞
𝑞=0 𝒄 𝒂 𝑨𝒂 𝒃𝒂 𝑞 𝑙−1 . Next, the forced response orig-
inated from the error 𝜖 is considered. Let 𝒆𝒙 (𝑘) be defined as the
estimation error 𝒙𝒂 (𝑘) − 𝒙̂ 𝒂 (𝑘). Suppose that the estimator gain vector
The plant to be controlled is assumed to be linear and stable in what 𝒍 is designed such that the matrix 𝑨𝒂 − 𝒍𝒄 𝒂 is stable. Then the predicted
follows. The state-space description of the plant is given as, error 𝒆̂ 𝒙 (𝑘 + 𝑖) is computed as,
{ ( )𝑖
𝒙𝑴 (𝑘 + 1) = 𝑨𝑴 𝒙𝑴 (𝑘) + 𝒃𝑴 𝒖 𝑢 (𝑘) + 𝒃𝑴 𝒅 𝑑 (𝑘) , 𝒆̂ 𝒙 (𝑘 + 𝑖) = 𝑨𝒂 − 𝒍𝒄 𝒂 𝒆𝒙 (𝑘) . (19)
(10)
𝑦𝑀 (𝑘) = 𝒄 𝑴 𝒙𝑴 (𝑘) ,
Therefore, the predicted 𝑖-step-ahead output error can be written as,
where 𝒙𝑴 ∈ R𝑛 denotes the state vector, 𝑢 ∈ R indicates the control ( )𝑖
𝜖̂ (𝑘 + 𝑖) = 𝒄 𝒂 𝑨𝒂 − 𝒍𝒄 𝒂 𝒆𝒙 (𝑘) . (20)
input, 𝑦𝑀 ∈ R represents the model output, and 𝑑 ∈ R signifies the
disturbance at the plant input. Here, it is assumed that 𝑨𝑴 , 𝒃𝑴 𝒖 and It can be seen from Eq. (20) and the definition of 𝒆𝒙 (𝑘) that 𝜖̂ (𝑘 + 𝑖)
𝒄 𝑴 satisfy the following condition to prevent the plant from having can be computed if 𝒙𝒂 (𝑘) is completely measurable, which enables
zeros at 1: the computation of the forced response due to 𝜖 (𝑘). However, only
( ) some of the state variables in 𝒙𝒂 (𝑘) are measurable in general, and
𝑨𝑴 − 𝑰 𝒃𝑴 𝒖
det ≠ 0. (11) it is impossible to apply the relation in Eq. (20) directly to compute
𝒄𝑴 0
the forced output. To get rid of the difficulty, the use of the following
( )
It is also assumed that 𝒄 𝑴 , 𝑨𝑴 is observable. The state-space descrip- approximation is suggested (Satoh et al., 2016):
tion of the disturbance 𝑑 is assumed to be given as, 𝜖̂ (𝑘 + 𝑖) ≃ 𝜖 (𝑘) = 𝒄 𝒂 𝒆𝒙 (𝑘) = 𝑦𝑃 (𝑘) − 𝒄 𝒂 𝒙̂ 𝒂 (𝑘) , ∀𝑖. (21)
( ) ( )( )
𝑑 (𝑘 + 1) 1 1 𝑑 (𝑘) It is clear from the right-hand side of Eq. (21) that only 𝑦𝑃 (𝑘), 𝒙̂ 𝒂 (𝑘)
= , (12)
𝛥𝑑 (𝑘 + 1) 0 1 𝛥𝑑 (𝑘) and 𝒄 𝒂 are required to utilize this approximation. Here, 𝑦𝑃 (𝑘) and 𝒙̂ 𝒂 (𝑘)
where 𝛥𝑑 (𝑘) ∶= 𝑑 (𝑘) − 𝑑 (𝑘 − 1). The following augmented system is are measurable, and 𝒄 𝒂 is a predefined constant vector. Thus, Eq. (21)
constructed using Eqs. (10) and (12): can be computed easily, and then used to compute the forced response
{ due to 𝜖.
𝒙𝒂 (𝑘 + 1) = 𝑨𝒂 𝒙𝒂 (𝑘) + 𝒃𝒂 𝑢 (𝑘) , Now, similarly to 𝑢 (𝑘 + 𝑖) in Eq. (4), 𝜖̂ (𝑘 + 𝑖) is assumed to be
(13)
𝑦𝑎 (𝑘) = 𝒄 𝒂 𝒙𝒂 (𝑘) , composed of the sum of weighted basis functions. However, only the
step basis function is used in this case from the assumption that 𝜖̂ (𝑘 + 𝑖)
where can be approximated as shown in Eq. (21). Hence 𝜖̂ (𝑘 + 𝑖) can be
⎛𝒙𝑴 (𝑘)⎞ ⎛𝑨𝑴 𝒃𝑴 𝒅 𝟎⎞ expressed as,
𝒙𝒂 (𝑘) ∶= ⎜ 𝑑 (𝑘) ⎟ ∈ R𝑛+2 , 𝑨𝒂 ∶= ⎜ 𝟎 1 1⎟ ∈ R(𝑛+2)×(𝑛+2) ,
⎜ ⎟ ⎜ ⎟ 𝜖̂ (𝑘 + 𝑖) = 𝜖 (𝑘) 𝑈𝐵1 (𝑖) , (22)
⎝ 𝛥𝑑 (𝑘) ⎠ ⎝ 𝟎 0 1⎠
(14) where 𝑈𝐵1 (𝑖) = 1. The forced response due to the error 𝜖 is expressed
⎛𝒃𝑴 𝒖 ⎞ ( )
𝒃𝒂 ∶= ⎜ 0 ⎟ ∈ R𝑛+2 , 𝒄 𝒂 ∶= 𝒄 𝑴 0 0 ∈ R1×(𝑛+2) . as,
⎜ ⎟
⎝ 0 ⎠ 𝑦̂𝐹2 (𝑘 + 𝑖) = 𝜖 (𝑘) 𝑦2𝐵1 (𝑖) , (23)
The following predictive state estimator for the augmented system ∑𝑖−1 𝑖−1−𝑞
where 𝑦2𝐵1 (𝑖) = 𝑞=0 𝒄 𝒂 𝑨𝒂 𝒍. The total forced response is therefore
defined in Eq. (13) is constructed: given as follows:
𝒙̂ 𝒂 (𝑘 + 1) = 𝑨𝒂 𝒙̂ 𝒂 (𝑘) + 𝒃𝒂 𝑢 (𝑘) + 𝒍𝜖 (𝑘) , (15) 𝑦̂𝐹 (𝑘 + 𝑖) = 𝑦̂𝐹1 (𝑘 + 𝑖) + 𝑦̂𝐹2 (𝑘 + 𝑖)
𝑛𝐵
where 𝜖 (𝑘) ∶= 𝑦𝑃 (𝑘) − 𝒄 𝒂 𝒙̂ 𝒂 (𝑘) = 𝑦𝑃 (𝑘) − 𝒄 𝑴 𝒙̂ 𝑴 (𝑘), 𝒍 ∈ R𝑛+2
is the ∑
= 𝜇𝑙 (𝑘) 𝑦1𝐵𝑙 (𝑖) + 𝜖 (𝑘) 𝑦2𝐵1 (𝑖) . (24)
estimator gain vector, and 𝑦𝑃 ∈ R is the plant output. Unlike the 𝑙=1
standard PFC in which Eq. (10) is utilized as the internal model, this ( )
The predicted output 𝑦̂𝑃 𝑘 + ℎ𝑗 in Eq. (2) is then written as,
estimator is used as the internal model in the PFC_ EBIM to compute
( ) ( ) ( )
the free and forced responses. 𝑦̂𝑃 𝑘 + ℎ𝑗 = 𝑦̂𝐹 𝑘 + ℎ𝑗 + 𝑦̂𝑈 𝐹 𝑘 + ℎ𝑗 . (25)

3
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

where 𝜂 > 0 is the leadscrew efficiency and is usually greater than


or equal to 0.9. It follows from Eqs. (27) and (28) that the torque
generated by the motor 𝜏 is written as,
1 { }
𝐽𝑀 𝜃̈ (𝑡) + 𝐾𝑅 𝐾𝑛 𝐾𝑅 𝜃 (𝑡) − 𝑥 (𝑡) = 𝜏 (𝑡) , (29)
𝜂
where 𝐽𝑀 is the moment of inertia. Here, the inertia of the screw is
considered as part of the motor’s inertia. The equation of motion of the
table is written as,
Fig. 2. Single-axis positioning system.
𝑀𝑡 𝑥̈ (𝑡) + 𝐷𝑡 𝑥̇ (𝑡) = 𝑃 , (30)

Table 1 where 𝑀𝑡 is the mass of the table, and 𝐷𝑡 is the viscous damping
Specifications of single-axis positioning system. coefficient. The following equation is obtained using Eqs. (27) and (30):
Parameter Specification
Ball screw lead 10 mm { }
𝑀𝑡 𝑥̈ (𝑡) + 𝐷𝑡 𝑥̇ (𝑡) + 𝐾𝑛 𝑥 (𝑡) − 𝐾𝑅 𝜃 (𝑡) = 0. (31)
Stroke 500 mm
Rail length 670 mm
If a DC servo amplifier is used in the torque control mode instead of
Rated speed 700 mm/s
manipulating the motor current directly, then the motor torque 𝜏 is
proportional to the command voltage applied to the motor driver 𝑒𝑎 ,
Table 2 and can be expressed as,
Specifications of DC motor.
Parameter Specification 𝜏 (𝑡) = 𝐾𝑆 𝑒𝑎 (𝑡) , (32)
Nominal voltage 24 V
where 𝐾𝑆 represents the proportional constant. Using Eqs. (29), (31)
Rated current 2.96 A
Rated torque 0.28 N m and (32), the transfer function from the command voltage 𝑒𝑎 to the
Rated speed 1810 r/min table displacement 𝑥 is computed as,
𝑋 (𝑠) 𝐾𝑆
= ( ) . (33)
𝐸𝑎 (𝑠) 𝐽𝑀 𝑀𝑡 4 𝐽𝑀 𝐷𝑡 3 𝐽𝑀 1 1
𝑠 + 𝑠 + + 𝐾𝑅 𝑀𝑡 𝑠2 + 𝐾𝑅 𝐷𝑡 𝑠
Then the optimal control law that minimizes the performance index 𝐾𝑛 𝐾𝑅 𝐾𝑛 𝐾𝑅 𝐾𝑅 𝜂 𝜂
given in Eq. (2) can be obtained by setting 𝜕𝐽 (𝑘) ∕𝜕𝝁 (𝑘) = 0. The If the stiffness 𝐾𝑛 is very high, then by taking the limit 𝐾𝑛 to infinity,
optimal control input is given as follows (Satoh et al., 2016): Eq. (33) can be rewritten as,
( )
𝑢 (𝑘) = 𝑘0 𝑐 (𝑘) − 𝑦𝑃 (𝑘) + 𝝂 𝑇𝒙 𝒙̂ 𝒂 (𝑘) + 𝑘𝜖 𝜖 (𝑘) , (26) 𝑋 (𝑠) 𝐾𝑆
= {( ) }. (34)
𝐸𝑎 (𝑠) 𝐽𝑀 1 1
where 𝑘0 ∈ R, 𝝂 𝒙 ∈ R𝑛+2 , and 𝑘𝜖 ∈ R are defined as shown in the 𝑠 + 𝐾𝑅 𝑀𝑡 𝑠 + 𝐾𝑅 𝐷𝑡
𝐾𝑅 𝜂 𝜂
Appendix.
Hence, it can be seen that the simplest model of a single-axis posi-
tioning system consists of a first-order lag system plus an integrator.
3. Plant description and modeling
Although Eq. (34) ignores vibration characteristics and, as a result,
the model becomes a one-inertia system, it is enough for our purpose
3.1. Plant description
in this paper. When 𝐾𝑛 goes to infinity, 𝑥 (𝑡) approaches 𝐾𝑅 𝜃 (𝑡), and
𝑋 (𝑠) = 𝐾𝑅 𝛩 (𝑠). Hence the transfer function between 𝑒𝑎 (𝑡) and 𝜃 (𝑡)
As a single-axis positioning system, the Monocarrier MCM08050H10K
becomes
(NSK, Ltd.) shown in Fig. 2 was used. As an actuator, an 80 W DC
𝛩 (𝑠) 𝐾𝑆
servomotor (Maxon Motor AG, F2260.885) equipped with an optical = {( ) }. (35)
𝐸𝑎 (𝑠) 1 1
encoder (1000 p/rev) was used, together with a DC servo amplifier 𝑠 𝐽𝑀 + 𝐾𝑅2 𝑀𝑡 𝑠 + 𝐾𝑅2 𝐷𝑡
(Maxon Motor AG, ADS 50/5). The allowable command voltage to the 𝜂 𝜂
amplifier was limited within ±10 V, which defined the hard constraints Let 𝐽 and 𝐷 be defined as follows:
on the control input. Tables 1 and 2 list the primary specifications of 1 2
𝐽 ∶= 𝐽𝑀 + 𝐾 𝑀, (36)
the single-axis positioning system and the DC motor. 𝜂 𝑅 𝑡
1 2
𝐷 ∶= 𝐾 𝐷. (37)
3.2. Plant modeling 𝜂 𝑅 𝑡
Then Eq. (35) can be expressed as,
This section is dedicated to the derivation of the dynamic model of 𝛩 (𝑠) 𝐾𝑆
a single-axis positioning system. = . (38)
𝐸𝑎 (𝑠) 𝑠 (𝐽 𝑠 + 𝐷)
It is presumed that a servomotor is connected to the feed screw of
Eq. (38) means that the position control of a single-axis positioning
a table drive system through a coupling. The force generated by the
system can be cast as the angle control of a motor, which is called a
screw is given by
semi-closed-loop control.
{ }
𝑃 = 𝐾𝑛 𝐾𝑅 𝜃 (𝑡) − 𝑥 (𝑡) , (27) The dominant disturbance in a single-axis positioning is friction, and
it can be regarded as a disturbance torque for the system in Eq. (38).
where 𝜃 is the motor angle, 𝑥 is the table position, 𝐾𝑛 is the stiffness Let 𝑇𝑑 be the disturbance torque other than the viscous friction torque,
between the table and the screw, and 𝐾𝑅 is the coefficient of trans- such as the Coulomb friction and breakaway torque. Also, let us assume
formation from the motor angle 𝜃 to the table position 𝑥. 𝐾𝑅 can be that 𝑇𝑑 is expressed using the equivalent voltage disturbance 𝑑 as
expressed as 𝐿∕2𝜋 where 𝐿 is the lead of the screw. The torque due to 𝑇𝑑 (𝑡) = 𝐾𝑆 𝑑 (𝑡). Then, the equation of motion of the motor is given
the force 𝑃 can be written as, by
1
𝑇𝑏 = 𝐾 𝑃, (28) 𝐽 𝜃̈ (𝑡) = 𝐾𝑆 𝑒𝑎 (𝑡) − 𝐷𝜃̇ (𝑡) + 𝑇𝑑 (𝑡) = 𝐾𝑆 𝑒𝑎 (𝑡) − 𝐷𝜃̇ (𝑡) + 𝐾𝑆 𝑑 (𝑡) . (39)
𝜂 𝑅

4
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

Fig. 3. Block diagram of industrial P–PI controller with anti-windup scheme.

Therefore, an equivalent state-space description of Eq. (34) with an


input disturbance term is given by

⎧ (𝑥̇ (𝑡)) (0 1
)(
𝑥1 (𝑡)
) (
0
) (
0
)
⎪ 1 = + 𝑒𝑎 (𝑡) + 𝑑 (𝑡) ,
⎪ 𝑥̇ 2 (𝑡) 0 −𝐷∕𝐽 𝑥2 (𝑡) 𝐾𝑆 ∕𝐽 𝐾𝑆 ∕𝐽
⎨ ( )
⎪ ( ) 𝑥1 (𝑡)
𝑦 (𝑡) = 1 0 ,
⎪ 𝑥2 (𝑡)

(40)
( )𝑇
where 𝑥1 𝑥2 ∶= (𝜃 𝜃) ̇ 𝑇 and 𝑦 ∶= 𝜃. A semi-closed-loop control system
is designed on the basis of Eq. (40) in this paper.
To use the plant model in Eq. (40), the values of the following
physical parameters of the single-axis positioning system need to be
known: 𝐽𝑀 , 𝑀𝑡 , 𝐷𝑡 and 𝜂. Unfortunately, they are not known exactly,
so 𝐽 in Eq. (36) and 𝐷 in Eq. (37) are identified directly, instead of 𝐽𝑀 , Fig. 4. Payload placed on slider.
𝑀𝑡 , 𝐷𝑡 and 𝜂. The standard system identification techniques (Ljung,
1999) can be applied to identify these parameters. However, care
should be exercised in the case of single-axis positioning systems when 4.2. Industrial P – PI controller
the system identification is conducted using the pseudo random binary
signal (PRBS) input, since the stroke of the slider is limited and the use
The industrial P – PI controller for a plant with the transfer function
of the PRBS input often moves the slider to the left or right end. To
of the form in Eq. (38) consists of a proportional–integral (PI) con-
avoid such an undesirable situation, the chirp signal was used as the
troller 𝐶PI in the inner velocity loop, and a proportional (P) controller
external input instead of the PRBS input. The prediction error method
in the MATLAB System Identification Toolbox (Mathworks, 2018) was 𝐶P in the outer position loop. These controllers are defined in the
utilized to identify the continuous-time model given in Eq. (40). The continuous-time domain as follows:
( )
identified values of these parameters are as follows: 𝐽 = 1.5219 × 𝐷
𝐶P (𝑠) ∶= 𝐾P , 𝐶PI (𝑠) ∶= 𝐾PI 1 + , (43)
10−4 kg m2 ; and, 𝐷 = 5.6201 × 10−4 Nm/(rad/s). The value of 𝐾𝑆 was 𝐽𝑠
excluded from the system identification since we were able to obtain where 𝐾P and 𝐾PI are the proportional gain and the integral gain,
the value experimentally as 𝐾𝑆 = 0.04 Nm/V. respectively. Fig. 3 contains a block diagram of the control system
The vectors and matrix appeared in the discrete-time state-space using the P – PI controller with the anti-windup scheme. It is clear from
model in Eq. (10) in this case are given as follows: Eq. (43) that the PI controller 𝐶PI (𝑠) cancels the nominal plant dynam-
( ) ( ) ics and forces the open-loop transfer function to become 𝐾𝑆 𝐾PI ∕𝐽 𝑠.
⎧ 1 0.9982 × 10−3 0.1314 × 10−3
⎪ 𝑨𝑴 = , 𝒃𝑴 𝒖 = , The closed-loop transfer function from the set-point to the plant output
⎨ 0 0.9963 0.2627 (41)
(
⎪𝒄 = 1 0 , ) becomes the following second-order system:
⎩ 𝑴 𝐾P 𝐾PI 𝐾𝑆
where the sampling period is taken to be 𝑇𝑠 = 1 ms. The vector 𝒃𝑴 𝒅 is 𝐽
𝑇 (𝑠) = (44)
the same as 𝒃𝑴 𝒖 . 𝐾PI 𝐾𝑆 𝐾 𝐾 𝐾
2
𝑠 + 𝑠 + P PI 𝑆
𝐽 𝐽
4. Position controllers for comparison Hence, the DC gain of the closed-loop system becomes 1, and the
steady-state position error is zero. According to the Routh–Hurwitz
4.1. Standard PFC stability criterion, the closed-loop system in Eq. (44) is stable in theory
as long as 𝐾P > 0 and 𝐾PI > 0, since 𝐾𝑆 > 0 and 𝐽 > 0. However,
The standard PFC differs from the PFC_ EBIM in that it does not
due to unmodeled dynamics, uncertainties, and nonlinear friction, the
use the augmented states and the state estimator. The plant model in
closed-loop system may become unstable, depending on the choice of
Eq. (10), the reference trajectory in Eq. (1), the performance index in
𝐾P and 𝐾PI . Therefore, fine tuning of these gains using the real plant
Eq. (2) are used, and the control law of the standard PFC is given as
will be necessary.
follows:
( )
𝑢 (𝑘) = 𝑘0 𝑐 (𝑘) − 𝑦𝑃 (𝑘) + 𝝂̃ 𝑇𝑥 𝒙𝑴 (𝑘) , (42) 5. Design and experimental evaluation
where 𝑘0 ∈ R and 𝝂̃ 𝑥 ∈ R𝑛 are defined as Appendix. In the
standard PFC, constrained control input produced by passing through In this section, the PFC_ EBIM is applied to the position-control
an appropriate limiter is supplied to the internal model to handle input of the single-axis positioning system, and its control performance is
constraints (Richalet & O’Donovan, 2009). compared with that of the standard PFC and the P – PI control system.

5
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

Table 3
Values of integral absolute error (IAE) and root of integral squared error (RISE) for
Experiment 1.
Control method Reference
Rectangular Sinusoidal Trapezoidal
IAE RISE IAE RISE IAE RISE
PFC_EBIM 13.8499 22.6866 21.2157 11.7927 10.8166 6.4490
PFC 27.7328 23.1315 32.1401 17.4888 26.9776 14.5877
P–PI 16.5331 22.8063 29.2778 16.2414 15.4489 9.0076

5.1. Controller design

The gains of the P – PI controller are tuned so as to minimize the


tracking error when the reference is rectangular, with an amplitude of
50 mm, and the same gains are used throughout the experiment. The
designed gains are 𝐾P = 13 and 𝐾PI = 0.6.
The estimator gain vector 𝒍 ∈ R4 required in the PFC_ EBIM is
decided by the pole-placement technique. All four estimator poles are
placed in the 𝑧-plane using Ackerman’s formula. The pole location is
initially chosen in the 𝑠-plane, and then it is transformed into the
𝑧-plane. In this design, the four poles are placed at −60 in the 𝑠-
plane, which is equivalently 0.9418 in the 𝑧-plane. Then the following
estimator gain vector is obtained:
( )𝑇
𝒍 = 0.2293 19.1461 2.9856 0.0438 . (45)

The number of the basis function required in the PFC_ EBIM and
PFC is taken as 𝑛𝐵 = 2, and the desired closed-loop time response
is set to be 𝑇CLTR = 0.1. The approaching ratio is taken as 𝛼 =
𝑒−3𝑇𝑠 ∕𝑇CLRT , and the following three coincidence points are used in the
( )
PFC_ EBIM and PFC: ℎ1 , ℎ2 , ℎ3 = (33, 50, 100), which corresponds to
( )
𝑇CLRT ∕3𝑇𝑠 , 𝑇CLRT ∕2𝑇𝑠 , 𝑇CLRT ∕𝑇𝑠 , respectively. Constraint handling in
the PFC_ EBIM is accomplished in the same way as the standard PFC.

5.2. Experimental evaluation

The following three types of the position reference are used:

• rectangular wave
• sinusoidal wave
• trapezoidal curve

The name of the trapezoidal curve comes from the fact that the shape
of the second derivative of the curve (equivalently, acceleration) con-
stitutes a trapezoid. The following experimental conditions are used:

Experiment 1 the amplitude of the reference is 50 mm; no payload is Fig. 5. Results of experiment 1 for rectangular reference (top: plant output, middle:
used tracking error, bottom: control input).

Experiment 2 the amplitude of the reference is 10 mm; payload is Table 4


used Values of integral absolute error (IAE) and root of integral squared error (RISE) for
Experiment 2.
In Experiment 2, the payload is placed on the slider of the single-axis Control method Reference
positioning system as shown in Fig. 4. The weight of the payload is Rectangular Sinusoidal Trapezoidal
4075.8 g including the fixing bolts.
IAE RISE IAE RISE IAE RISE
PFC_EBIM 1.4030 3.0837 4.2436 2.3566 2.1632 1.2890
Experiment 1
Fig. 5 shows the reference and the plant output (top), the tracking PFC 6.8536 4.7820 17.5949 9.6204 16.9859 9.7292
error (middle) and the control input (bottom) of each control scheme P–PI 1.8952 3.3092 6.2292 3.4337 3.3423 1.9327
for the rectangular reference. It is clear from Fig. 5 that steady-state
error arises in the case of the standard PFC. This is because the PFC does
not actively estimate or reject the disturbance, and furthermore, the
system. Nonetheless, a small overshoot appears, and the settling time
integral action of the PFC controller is lost due to pole-zero cancellation
of the P – PI control system is rather long. On the other hand, the
since the single-axis positioning system is an integrating process. Both
the PFC_ EBIM and the P – PI succeed in tracking the reference without PFC_ EBIM tracks the reference without overshoot, and the settling time
any offset. As shown in Fig. 5, the control input for this reference is far shorter than that of the P – PI controller. It can be observed
reaches ±10 V, and is thus clipped by the limiter. In this case, the anti- from Fig. 5 that a small oscillation appears in the control input of the
windup scheme shown in Fig. 3 works properly in the P – PI control PFC_ EBIM at the steady-state. This phenomenon seems to be caused by

6
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

Fig. 6. Results of experiment 1 for sinusoidal reference (top: plant output, middle: Fig. 7. Results of experiment 1 for trapezoidal reference (top: plant output, middle:
tracking error, bottom: control input). tracking error, bottom: control input).

the controller tying to further reduce the position error under friction and P – PI. The values of the integral absolute error (IAE) and the root
in the same way as the hunting or limit cycles induced by an integral of the integral squared error (RISE) are computed from the tracking
controller. error data and shown in Table 3.
Fig. 6 shows the experimental results for the sinusoidal reference.
The plant output results appear similar. However, taking a look at the Experiment 2
tracking error, it can be seen that the PFC_ EBIM has better tracking A payload of 4075.8 g is introduced on the slider to investigate the
performance than the PFC or the P – PI. When the reference is close to robustness against model variation and disturbance rejection property
the extreme value, the PFC_ EBIM and P – PI achieve similar tracking of each control scheme. In addition, the amplitude of the reference
performance. Still, the PFC_ EBIM is slightly better than the P – PI (see is reduced from 50 mm to 10 mm. As a result, the effect of stiction
the close-up in Fig. 6). becomes relatively large, and position control in this experiment is
Fig. 7 shows the experimental results for the trapezoidal curve more difficult than that in Experiment 1.
reference. The tracking performance for the trapezoidal reference is Figs. 8–10 show the experimental results. In all cases, the PFC_ EBIM
more susceptible to stiction than that for the sinusoidal reference since achieves the fastest response and the smallest tracking error, even in
the rate of change of the trapezoidal curve is small, and hence the the presence of model variation and large stiction. On the other hand,
control input also becomes small, especially when the velocity is close performance degradation of other two control strategies is apparent.
to zero. Even for the trapezoidal reference, it can be seen from Fig. 7 For example, the response of the P – PI control for the rectangular
that the PFC_ EBIM achieves better tracking performance than the PFC reference is slow and has only a little overshoot, which is in contrast to

7
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

Fig. 8. Results of experiment 2 for rectangular reference (top: plant output, middle: Fig. 9. Results of experiment 2 for sinusoidal reference (top: plant output, middle:
tracking error, bottom: control input). tracking error, bottom: control input).

the result in Experiment 1. The responses of the standard PFC are far results showed that the PFC_ EBIM works well and achieved the best
worse than those in Experiment 1. It can be noticed that a large offset tracking performance of the three control strategies.
arises even for the sinusoidal and trapezoidal references, as well as for It turned out that small oscillation appears in the control input of
the rectangular reference. Again, oscillation in the control input of the the PFC_ EBIM especially for the case of the rectangular reference at the
PFC_ EBIM for the rectangular reference can be observed at the steady- steady-state. This issue will be addressed in future work.
state (see Fig. 8). The values of the integral absolute error (IAE) and
the root of the integral squared error (RISE) are presented in Table 4.
Acknowledgments
6. Conclusions

This work was supported by Japan Society for the Promotion of


An application of the predictive functional control with a state
estimator-based internal model (PFC_ EBIM) to a single-axis positioning Science under a Grant-in-Aid for Scientific Research (C) JP17K06233.
system was presented, and its practicability was validated experimen-
tally in this paper. The tracking performance of the PFC_ EBIM was
Conflict of interest
experimentally compared with those of the standard PFC and industrial
P – PI controller using the three types of reference, including a rect-
angular, sinusoidal, and modified trapezoidal curve. The experimental None declared.

8
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

( ) ( ( ) ( ))𝑇
where 𝒚 𝟏𝑩 ℎ𝑗 = 𝑦1𝐵1 ℎ𝑗 ⋯ 𝑦1𝐵𝑛 ℎ𝑗 ∈ R𝑛𝐵 , and 𝒖𝑩 (0) =
𝐵
(1 0 ⋯ 0)𝑇 ∈ Z𝑛𝐵 .
The vector 𝝂̃ 𝑥 in the standard PFC is defined as follows:
( ) 𝑇

⎛ 𝒄 𝑴 𝑨𝑴1 − 𝑰 ⎞
⎜ ⎟
𝝂̃ 𝑥 = − ⎜ ⋮ 𝝂̃ ∈ R𝑛 ,
( ℎ )⎟
⎜ 𝑛ℎ ⎟
⎝𝒄 𝑴 𝑨𝑴 − 𝑰 ⎠
where
(𝑛 )−1
( ( ) ( ))𝑇 ∑ ℎ
( ) ( )𝑇
𝝂̃ = 𝒚 𝑩 ℎ1 ⋯ 𝒚 𝑩 ℎ 𝑛ℎ 𝒚 𝑩 ℎ𝑗 𝒚 𝑩 ℎ𝑗 𝒖𝑩 (0) ∈ R𝑛ℎ
𝑗=1

( ) ( ( ) ( ))𝑇
and 𝒚 𝑩 ℎ𝑗 = 𝑦𝐵1 ℎ𝑗 ⋯ 𝑦𝐵𝑛 ℎ𝑗 ∈ R𝑛𝐵 and
𝐵


𝑖−1
𝑦𝐵𝑙 (𝑖) = 𝒄 𝑴 𝑨𝑖−1−𝑞
𝑴
𝒃𝑴 𝒖 𝑞 𝑙−1 .
𝑞=0

The vector 𝒖𝑩 (0) and the scalar 𝑘0 in the standard PFC are the same as
those in the PFC_ EBIM.

References

Abdelghani-Idrissi, M. A., Arbaoui, M. A., Estel, L., & Richalet, J. (2001). Predictive
functional control of a counter current heat exchanger using convexity property.
Chemical Engineering and Processing: Process Intensification, 40(5), 449–457.
Anzehaee, M. M., Behnam, B., & Hajihosseini, P. (2018). Augmenting armarkov-pfc
predictive controller with pid-type iii to improve boost converter operation. Control
Engineering Practice, 79, 65–77.
Abu el Ata-Doss, S., Fiani, P., & Richalet, J. (1991). Handling input and state constraints
in predictive functional control. In Proceedings of the 30th conference on decision and
control (pp. 985–990).
Bigdeli, N., & Haeri, M. (2009). Predictive functional control for active queue
management in congested tcp/ip networks. ISA Transactions, 48(1), 107–121.
Bouhenchir, H., Cabassud, M., & Lann, M. V. L. (2006). Predictive functional control
for the temperature control of a chemical batch reactor. Computers and Chemical
Engineering, 30, 1141–1154.
Camacho, E. F., & Bordons, C. (2007). Model predictive control (2nd ed.).
Springer-Verlag.
Dieulot, J. Y., Benhammi, T., Colas, F., & Barre, P. J. (2008). Composite predictive
functional control strategies, application to positioning axes. International Journal
of Computers Communications & Control, 3(1), 41–50.
Dovžan, D., & Škrjanc, I. (2010). Predictive functional control based on an adaptive
fuzzy model of a hybrid semi-batch reactor. Control Engineering Practice, 18,
979–989.
Franklin, G. F., Powell, J. D., & Workman, M. L. (1998). Digital control of dynamic
systems (3rd ed.). Addison-Wesley.
Friebel, T., Zabet, K., Haber, R., & Jelali, M. (2015). Predictive functional control
of tandem cold metal rolling. In Proceedings of the IEEE conference on control
applications (pp. 324–329).
Fig. 10. Results of experiment 2 for trapezoidal reference (top: plant output, middle: Kuntze, H. B., Jacubasch, A., Richalet, J., & Arber, C. (1986). On the predictive
tracking error, bottom: control input). functional control of an elastic industrial robot. In Proceedings of the 25th conference
on decision and control (pp. 1877–1881).
Lepetič, M., Škrjanc, I., Chiacchiarini, H. G., & Matko, D. (2003). Predictive functional
control based on fuzzy model: magnetic suspension system case study. Artificial
Appendix
Intelligence, 16, 425–430.
Liu, H., & Li, S. (2012). Speed control for pmsm servo system using predictive functional
The vectors and matrices in the PFC_ EBIM are defined as follows: control and extended state observer. IEEE Transactions on Industrial Electronics,
( ) 𝑇 59(2), 454–460.

⎛ 1 − 𝛼 ℎ1 ⎞ ⎛ 𝒄 𝒂 𝑨𝒂1 − 𝑰 ⎞ Ljung, L. (1999). System identification: Theory for the user (2nd ed.). Prentice Hall.
⎜ ⎟
𝑘0 ∶= 𝝂 𝑇 ⎜ ⋮ ⎟ ∈ R, 𝝂 𝒙 ∶= − ⎜ 𝑛+2 Maciejowski, J. M. (2002). Predictive control with constraints. Pearson Education.
⋮ ⎟ 𝝂∈R ,
⎜ ⎟ ( ) Mathworks (2018). System identification toolbox.
ℎ𝑛
⎝1 − 𝛼 ⎠ ℎ ⎜ ℎ𝑛
ℎ ⎟
⎝𝒄 𝒂 𝑨𝒂 − 𝑰 ⎠ Matsubara, A. (2008). Design and control of precision positioning and feed drive systems
( ) (in Japanese). Tokyo: Morikita Publishing.
⎛ 𝑦2𝐵 ℎ1 ⎞ Muñoz, W. A. P., Sellier, A. G., & Castro, S. G. (2018). The predictive functional control
⎜ 1
⋮ )⎟⎟ ∈ R.
𝑘𝜖 ∶= −𝝂 𝑇 ⎜ ( and the management of constraints in guanay ii autonomous underwater vehicle
⎜𝑦2𝐵 ℎ𝑛 ⎟ actuators. IEEE Access, 6, 22353–22367.
⎝ 1 ℎ ⎠ Nagase, J., Hamada, K., Satoh, T., Saga, N., & Suzumori, K. (2013). Comparison between
pfc and pid control system for tendon-driven balloon actuator. In Proceedings of the
Therein,
39th annual conference of the IEEE industrial electronics society (pp. 3396–3401).
(𝑛 )−1
( ( ) ( ))𝑇 ∑ ℎ
( ) ( )𝑇 Nagase, J., Satoh, T., Saga, N., & Suzumori, K. (2013). Predictive functional control
𝝂 ∶= 𝒚 𝟏𝑩 ℎ1 ⋯ 𝒚 𝟏𝑩 ℎ𝑛ℎ 𝒚 𝟏𝑩 ℎ𝑗 𝒚 𝟏𝑩 ℎ𝑗 of tendon-driven actuator using pneumatic balloon. Journal of Advanced Mechanical
𝑗=1 Design, Systems, and Manufacturing, 7(4), 752–762.
Qin, S. J., & Badgwell, T. A. (2003). A survey of industrial model predictive control
𝒖𝑩 (0) ∈ R𝑛ℎ , technology. Control Engineering Practice, 11(7), 733–764.

9
T. Satoh, N. Saito, J. Nagase et al. Control Engineering Practice 86 (2019) 1–10

Richalet, J. (1993). Industrial applications of model based predictive control. Satoh, T., Saito, N., Nagase, J., & Saga, N. (2016). Predictive functional control using
Automatica, 29(5), 1251–1274. state estimator-based internal model for ramp disturbance rejection. International
Richalet, J., Abu el Ata-Doss, S., Arber, C., Kuntze, H. B., Jacubasch, A., & Schill, W. Journal of Automation and Control, 10(3), 267–285.
(1987). Predictive functional control - application to fast and accurate robot. In Satoh, T., Saito, N., Nagase, J., & Saga, N. (2017). An application of predictive
Proceedings of the IFAC 10th world congress (pp. 251–258). functional control with a state observer-type internal model. In Proceedings of the
Richalet, J., Darure, T., & Mallet, J. (2014). Predictive functional control of counter IEEE 24th international conference on mechatronics and machine vision in practice (pp.
current heat exchangers. In Proceedings of the IFAC 19th world congress (pp. 266–271).
5345–5350). Satoh, T., Saito, N., & Saga, N. (2010). Predictive functional control with distur-
Richalet, J., & O’Donovan, D. (2009). Predictive functional control: principles and industrial bance observer for pneumatic artificial muscle actuator. In Proceeding of the 1st
applications. London, England: Springer-Verlag. international conference on applied bionics and biomechanics.
Rossiter, J. A. (2002). Predictive functional control: more than one way to prestabilise. Vivas, A., & Poignet, P. (2005). Model based predictive control of a fully parallel robot.
In Proceedings of the IFAC 15th world congress (pp. 289–294). Control Engineering Practice, 13(7), 863–874.
Rossiter, J. A. (2004). Model-based predictive control - a practical approach. CRC press. Zabet, K., & Haber, R. (2017). Robust tuning of pfc (predictive functional control) based
Rossiter, J. A., Haber, R., & Zabet, K. (2016). Pole-placement predictive functional on first- and aperiodic second-order plus time delay models. Journal of Process
control for over-damped systems with real poles. ISA Transactions, 61, 229–239. Control, 53, 25–37.
Satoh, T., Kaneko, K., & Saito, N. (2012). Improving tracking performance of predictive Zhang, R., Xue, A., Wang, S., & Zhang, J. (2012). An improved state-space model
functional control using disturbance observer and its application to table drive structure and a corresponding predictive functional control design with improved
systems. International Journal of Computers Communications & Control, 7(3), 550–564. control performance. International Journal of Control, 85(8), 1146–1161.
Zhang, B., Yang, W., Zong, H., Wu, Z., & Zhang, W. (2011). A novel predictive control
algorithm and robust stability criteria for integrating processes. ISA Transactions,
50, 454–460.

10

Vous aimerez peut-être aussi