You are on page 1of 13

Control Engineering Practice 59 (2017) 64–76

Contents lists available at ScienceDirect

Control Engineering Practice


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

Experimental sensorless control for IPMSM by using integral backstepping


strategy and adaptive high gain observer
crossmark

M.A. Hamidaa, , J. de Leonb, A. Glumineauc
a
Department of Electronics, Electrical Engineering Laboratory, Ouargla University, Ouargla, Algeria
b
Universidad Autonoma de Nuevo Leon, Facultad de Ingenieria Mecanica y Electrica, Mexico
c
LUNAM Université, Ecole Centrale de Nantes, IRCCyN, UMR CNRS 6597, 44321 Nantes, France

A R T I C L E I N F O A BS T RAC T

Keywords: In this paper a robust sensorless control for an Interior Permanent Magnet Synchronous Motor (IPMSM) is
Nonlinear sensorless control designed. The proposed control strategy uses a backstepping controller, whose robustness is improved by using
Interior Permanent Magnet Synchronous integral actions added at each step of the original algorithm, and by a Maximum-Torque-Per-Ampere strategy
Motor (IPMSM) (MTPA) to improve its energy efficient operation. Furthermore, to implement this controller in the framework of
Nonlinear observer
the mechanical sensorless control from the only measurements of the currents and voltages, an adaptive
Backstepping
interconnected high gain observer is developed for estimating the rotor speed, the position and the load torque.
Experimental results
Moreover, sufficient conditions are given to ensure the practical stability of the Observer-Controller system even
if bounded uncertainties occur. Finally, the performance and the effectiveness of the designed method are tested
experimentally throw a significant benchmark including different speed references and with significant
robustness tests. A comparative evaluation of the computational effort of our scheme with respect to classical
motor control is given.

1. Introduction 2013b). The one based on the high-frequency injection signal brings
high-frequency noise. In most cases, this noise happens to be at
The permanent-magnet synchronous motor (PMSM) has been frequencies higher that the bandwidth of the regulators which can be
widely used in the industry for variable speed applications due to its avoided by using analogical filter (Qiao et al., 2013) but has to be below
high performance reliability and its power density. Owing to the the PWM frequency.
progress in the permanent magnet materials, micro and power On the other hand, the observer design based on the sliding mode
electronics, fast digital signal processors and modern control technol- techniques is widely used (Qiao et al., 2013) thanks to its performance
ogies, the permanent magnet synchronous machines have become and its robustness (Ren, Liu, Wang, & Liu, 2015). Nevertheless, the
more widespread in the industrial applications, for example: auto- chattering phenomenon presents the principal drawback of this kind of
mobiles, robotics, aeronautics and aerospace domains. observers (Liu, Laghrouche et al., 2014; Wang, Li, Zhang, Yu, & Xu,
Furthermore, high-performance controls require continuous mea- 2013). Furthermore, due to its simplicity, the back electromotive force
surement of rotor position and speed. Then, due to the cost and is widely used for the rotor speed estimation (Lu, Lei, & Blaabjerg,
reliability constraints (Betin et al., 2014; Delpoux, Bodson, & Floquet, 2013a).
2014; Liu & Zhu, 2014), the elimination of shaft position sensors is Regarding the observation problem of PMSM, the observer design
desirable for industrial applications (Kshirsagar et al., 2012; Raggl, of a PMSM using the methods based on the back EMF is more sensitive
Warberger, Nussbaumer, Burger, & Kolar, 2009). to parametric deviations. More precisely, the information of position is
In the literature, several methods have been proposed to estimate found in the back EMF and in the inductances can be used to estimate
the rotor speed in order to achieve high performance sensorless the position and the speed. Nevertheless, the electromotive force
control. These methods are arranged in two principal classes of the becomes too small at very low speeds which make the speed estimation
non model-based approaches: the high-frequency injection (Kim, Ha, very difficult. To address this problem, the extended electromotive
& Sul, 2012; Liu & Zhu, 2014) and artificial intelligence methods force is proposed as an alternative strategy (Wang, Zhan, Zhang, Gui,
(Accetta, Cirrincione, Pucci, & Vitale, 2012; Lu, Lei, & Blaabjerg, & Xu, 2014). The rotor speed can be also obtained from the flux


Corresponding author.
E-mail addresses: assaad.hamida@hotmail.fr, hamida.assaad@univ-ouargla.dz (M.A. Hamida).

http://dx.doi.org/10.1016/j.conengprac.2016.11.012
Received 22 September 2015; Received in revised form 17 November 2016; Accepted 18 November 2016
0967-0661/ © 2016 Elsevier Ltd. All rights reserved.
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

linkage estimation (Xu, Zhang, & Liu, 2013). This method is easy to
implement. However, at low speeds is not sufficient to obtain a good
estimation when the parameters are not exactly known (Hamida,
Glumineau, & de Leon, 2012).
For most of the observers, the parameter uncertainties (Niapour,
Tabarraie, & Feyzi, 2014) and an unknown load torque represent one
of the main difficulties for the IPMSM observation and control.
Thus, in this paper, a nonlinear observer is implemented experi-
mentally to estimate on-line the rotor resistance and the load torque.
Then this estimation will be used for the sensorless control of the
IPMSM, which allows to increase the robustness of the proposed
controller-observer scheme under parameter variations.
It is worth mentioning that the most of the sensorless control
solutions proposed in the literature do not take into account this
difficulty. From this point of view, nonlinear robust controls as sliding
mode (Alwi & Edwards, 2014) and backstepping (Escareno,
Rakotondrabe, & Habineza, 2015) could be more efficient. In
Hamida et al. (2012), the classical backstepping control is robustified
by adding tracking errors integral actions. This controller has been
tested only in simulation. Moreover, in order to guarantee an energy
efficient operation of the IPMSM, this controller is combined with a
MTPA strategy maximize the output torque. Fig. 1. PMSM rotor permanent magnets layout: (a) Surface permanent magnets, (b)
Inset permanent magnets, (c) Interior permanent magnets (d) Flux concentrating.

1.1. Contribution 2.1.1. Classification of Permanent Magnet Synchronous Motors


(PMSM)
In this paper a nonlinear sensorless control for IPMSM is devel- The physical characteristics of the PMSM are associated with its
oped. A robust nonlinear control strategy namely integral MTPA rotor and stator structures.
backstepping controller is proposed to achieve the control objectives. The stator is composed of a three-phase winding such that the
Furthermore, to implement this controller, a high gain interconnected Electromotive Forces (EMF) are generated by the rotation of the rotor
observer is designed for estimating the rotor position, the rotor speed field. The rotor incorporates permanent magnets to produce a magnetic
at wide speed range, which allows to replace the mechanical sensors. field. Regarding the winding, the permanent magnets have the
Moreover, the proposed observer is also used to estimate the load advantage to eliminate the brushes, the rotor losses and the need of
torque and stator resistance, in order to improve the robustness of the a controlled DC source to provide the excitation current. However, the
developed strategy. The stability of the proposed observer-controller amplitude of the rotor flux is constant.
scheme is ensured using Lyapunov approach. Finally, the designed On the other hand, there exist several ways to place the magnets in
sensorless control is tested experimentally on a significant industrial the rotor (see Fig. 1).
sensorless benchmark. These experiments results are the main con- Following the magnets position, the PMSM can be classified into
tribution of this paper, where the good performance and the robustness four major types: (a) Surface Mounted magnets type (Fig. 1a), (b) Inset
of the control strategy for the sensorless IPMSM is shown, and its magnets type (Fig. 1b), (c) Interior magnets type (Fig. 1c), (d) Flux
practical implementation in usual low-cost industrial drives is dis- concentrating type (Fig. 1d).
cussed. However, surface permanent magnet synchronous motors and
interior permanent magnet synchronous motors are the most used in
the industry. Particularly, the synchronous machines with sinusoidal
1.2. Paper structure EMF are classified in two subcategories in terms of magnets position.

The structure of the paper is as follows. In Section 2, the Permanent 1. Non-Salient Poles: the magnets are located in the rotor surface
Magnet Synchronous Motor (PMSM) and the nonlinear mathematical (Fig. 1a): Surface Permanent Magnet Synchronous Motor (SPMSM)
model of the Interior Permanent Magnet Synchronous Motor(IPMSM) 2. Salient Poles: the magnets are buried into the rotor (Fig. 1c and
are introduced. Furthermore, before to design an observer, the d): Interior Permanent Magnet Synchronous Motor (IPMSM).
observability of the system is firstly recalled in Section 3 and then, a
high gain adaptive observer is developed in Section 4. The convergence 2.1.2. Model of the IPMSM
of the proposed observer under parameter uncertainties is studied in The mathematical model of the IPMSM in the (d , q ) reference frame
Section 5. Furthermore in Section 6, a robust Backstepping Controller for sinusoidal stator current excitation is given by Pillay and Krishnan
is introduced. Next, the stability of the observer-controller scheme is (1989):
proved in Section 7. Experiments are carried out showing the effec- did R Lq 1 diq R L 1
= − s id + p Ωiq + vd = − s iq − p d Ωid − p ϕf Ω
tiveness of the sensorless control strategy and reported in Section 8. dt Ld Ld L d dt Lq Lq Lq
Finally, some conclusions are given in Section 9. 1 dΩ p f p 1 dθ
+ vq = (L d − Lq ) id iq − v Ω + ϕf iq − Tl =Ω
Lq dt J J J J dt (1)
2. Problem statement with

2.1. Permanent magnet synchronous motor Rs stator resistance.


L d , Lq dq-axis inductances.
In the sequel, we introduce the most important characteristics of ϕf permanent-magnet flux linkage.
the PMSM.

65
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

id , iq stator currents. ⎡ iq ⎤ ⎡ Lq ⎤
0 −L ⎥ ⎢ 0 p Ld iq ⎥
vd , vq stator voltages. A1 (·) = ⎢ q , A2 (·) = ⎢ ,
⎢⎣ ⎥ fv ⎥
Ω rotor mechanical speed. 0 0 ⎦ ⎢⎣ 0 − J ⎥⎦
θ rotor angular position. ⎡ ϕf ⎤
L
− p Ld Ωid − p L Ω ⎥ ⎡ 0 ⎤
J moment of inertia. g1 (·) = ⎢ q q , Φ = ⎢ 1⎥
fv viscous friction coefficient. ⎢ ⎥ ⎢⎣− J ⎥⎦
⎣ 0 ⎦
p number of pole pairs.
Tl load torque. ⎡ R
− L s id ⎤ ⎡1⎤ ⎡1⎤
g2 (·) = ⎢⎢ p d ⎥,
⎥ Φ1 = ⎢ Lq ⎥ , Φ 2 = ⎢ Ld ⎥,
p ⎢⎣ 0 ⎥⎦ ⎢⎣ 0 ⎥⎦
⎣ J ϕf iq + J (L d − Lq ) iq id ⎦
Definition 1. The IPMSM physical operation domain + is defined by
the set of values C1 = C2 = [10],

+ = {X ∈ 5| id ≤ Idmax , iq ≤ Iqmax , Ω ≤ Ω max , Tl ≤ Tlmax , Rs ≤ Rsmax } with X1 = (iq Rs )T , X2 = (id Ω )T are the states, u = (vd vq )T is the input, and
y = (id iq )T is the output of the IPMSM model. Furthermore, g1 and g2
are the interconnection terms.
with X = [id iq Ω Tl Rs]T and Idmax, Iqmax, Ωmax, Tlmax and Rsmax are
For each subsystem, the state of the other subsystem is supposed
the actual maximum values for currents, speed, load torque and stator
available. In the sequel, a nonlinear observer will be developed for each
resistance, respectively.
subsystem. Then, the nonlinear observer for the IPMSM is composed
from the interconnection of the two observers.
3. IPM synchronous motor observability analysis
Furthermore, the convergence of the interconnected observer for
the IPMSM model, represented by (2),(3) and (4), is ensured provided
In this section, the observability property of the IPM synchronous
that the following properties are satisfied (see Besancon & Hammouri,
machine is analyzed. This study is made to decide if it is possible to
1998 for details):
compute the estimates of the rotor position, the rotor speed, the load
torque and also the stator resistance parameter based only on the
1. The matrices A1 et A2 are only functions of the measurement y1 = iq
measured stator voltages and the stator currents (for more details see
and the observability property of each subsystem is ensured if the
Hamida, Leon, & Glumineau, 2014).
signal y1 = iq has sufficient information to ensure the observability
From the mathematical model of the IPMSM in the d-q reference
property.
frame (1), based on the observability analysis given in Hamida et al.
2. The term g1 (X2 ) is globally Lipschitz w.r.t. X2.
(2014), and for the IPM Synchronous Motor, i.e. (Lq ≠ L d ), it follows
3. The term g2 (X1, y2 ) is globally Lipschitz w.r.t. X1 and uniformly w.r.t.
that the rotor speed and rotor position of the IPMSM are observable for
y2.
all rotor speed values, including rotor speed equal zero. For the SPM
Synchronbous Motor, i.e. (Ld=Lq), the rotor speed and rotor position
Then, a nonlinear adaptive interconnected observer for (3) and (4) is
of the IPMSM are observable, if the rotor speed is different from zero.
given by:
This can be summarized as follows.
⎧ Z˙ = A (y ) Z + g (Z ) + Φ (u ) + S −1 C T (y − yl )
1. The Surface PMSM (SPMSM) is not observable at zero speed. ⎪ 1 1 1 1 1 2 1 1 1 1 1
O1: ⎨ S˙1 = −ρ1 S1 − A1T (y1) S1 − S1 A1 (y2 ) + C1T C1
2. The rotor speed and the rotor position of the Interior PMSM ⎪
(IPMSM) are observable at all speed ranges including at zero speed. ⎩ yl1 = C1 Z1 (5)
3. The stator resistance and the load torque of the Interior PMSM are
⎧ Z˙ = A2 (y1) Z 2 + g2 (Z1, y2 ) + Φ 2 (u ) + ΦTll + KC1T (y1 − yl1)
observable at all speed ranges except at standstill (the currents and ⎪ 2
the voltages are null). This particular case can be easily detected by ⎪ + (ϖΛS3−1 ΛT C2T + ΓS2−1 C2T )(y2 − yl2 )

the stator measurements. ⎪ Tl˙l = ϖS3−1 ΛT C2T (y2 − yl2 ) + B (Z1)(y1 − yl1) + B1 (Z1)(y2 − yl2 )

O2 : ⎨ S˙2 = − ρ2 S2 − A2T (y1) S2 − S2 A2 (y1) + C2T C2
4. Nonlinear interconnected observers design for IPMSM ⎪
⎪ S˙3 = − ρ3 S3 + ΛT C2T C2 Λ
⎪ ˙
In this section, a nonlinear interconnected observer is designed for ⎪ Λ = {A2 (y1) − ΓS2−1 C2T C2} Λ + Φ
⎪ yl
estimating simultaneously the rotor speed, rotor position, load torque ⎩ 2 = C2 Z 2 (6)
and the stator resistance. For this purpose the mathematical model of
the PMSM given in (1) can be represented in two subsystems, where T T
the dynamics of the load torque and the stator resistance are included with Z1 = [i^q mRs ] and Z 2 = [i^d Ω l ] are the estimated state variables
in each subsystem. These dynamics are assumed to be piece-wise respectively of X1 and X2. ϖ is a positive constant, which represents the
functions, which can be expressed as follows: gain of the load torque observer. Λ is the solution of dynamical system,
which works as a filter system. The ρi are positive constants, for
T˙l = KTl R˙s = K Rs, (2) i = 1, 2, 3. Furthermore, S1 and S2 are symmetric positive definite
with KTl and K Rs are positive constants. matrices (Besancon & Hammouri, 1996), with S3 (0) > 0 ,
B (Z1) = k J Φf i^q , B1 (Z1) = k J (L d − Lq ) i^q
p p
From (1) and (2), the IPMSM mathematical model can be decom-
posed into two interconnected subsystems (3) and (4) (Hamida et al., ⎡ − kc1 ⎤ ⎡ ⎤
2012): K = ⎢− k ⎥ , Γ = ⎢1 0 ⎥
⎢⎣ c2 ⎥⎦ ⎣0 α⎦
⎧ X˙1 = A1 (y ) X1 + g (X2 ) + Φ1 u
Σ1: ⎨ 1 1
⎩ y1 = C1 X1 (3)
with k, kc1, kc2 and α and are positive constants.
⎧ X˙2 = A2 (y ) X2 + g (X1, y ) + Φ 2 u + ΦTl From the structure of the proposed observer, the first subsystem (5)
Σ 2: ⎨ 1 2 2
represents an observer used to estimate the current iq and the stator
⎩ y2 = C2 X2 (4)
resistance Rs. This is a copy of the system plus a term depending of the
where error between the measured and estimated outputs. Furthermore, the

66
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

gain of this observer is given by S1−1 C1T , which depends on the solution ⎡ ⎛L ⎞ ⎤
of a matrix differential equation of S1 (differential Ricatti equation). ⎢ 0 pΔ ⎜ q ⎟ iq ⎥
ΔA2 (·) = ⎢ ⎝ Ld ⎠ ⎥,
The solution of the differential equation on S1 uses the following lemma ⎢ f ⎥
(see proof in Besancon, De Leon, & Huerta, 2006), which will be used ⎢⎣ 0 Δ ( Jv ) ⎥⎦
to ensure the convergence of the developed observer. ⎡ ⎛1⎞ ⎤
⎢ − Rs Δ ⎜ L ⎟ id ⎥
Lemma 1. Assume that the input v is regularly persistent for a given Δg2 (·) = ⎢ ⎝ ⎠d ⎥.
state affine system and consider the following Lyapunov differential ⎢ 1 1 ⎥
⎢⎣ − pΔ ( J
) ϕ i i
f d q − pΔ ( J
)(Δ L d − ΔL q q⎥
) i ⎦
equation
The Δ(.) are bounded uncertain terms with respect to the uncertain
S˙ (t ) = −θS (t ) − AT (v (t )) S (t ) − S (t ) A (v (t )) + C T C.
parameters, then there exist positive constant ϱi > 0 , for i = 1, …, 4 ,
such that:
with S (0) > 0 , then ∃ θ0, ∀ θ ≥ θ0, ∃ α > 0, β > 0, t0 > 0 :
ΔA1 ≤ ϱ1, ΔA2 ≤ ϱ2 , Δg1 ≤ ϱ3, Δg2 ≤ ϱ4 .
∀ t0, αI ≤ S (t ) ≤ βI
The estimation errors are defined as:
were I is the identity matrix.
From Lemma 1, it is clear that for subsystem (5), v = y1 and ϵ1 = X1 − Z1, ϵ′2 = X2 − Z 2, ϵ3 = Tl − Tll . (10)
S (t ) = S1. From Eqs. (5)–(6) and (8)–(9). Using the same idea as in Zhang
For the second subsystem (6), an observer is designed to estimate (2002), we introduce the following transformation ϵ2 = ϵ′2 − Λϵ3. The
the current id, the rotor speed Ω, and the load torque Tl. The current id estimation errors dynamics can be expressed as:
and the rotor speed Ω are estimated by using the observer, whose gain
is split into two parts. The first part is the gain ΓS2−1 C2T obtained from ϵ̇1 = [A1 (y1) − S1−1 C1T C1]ϵ1 + g1 (X2 ) − g1 (Z 2 ) + ΔA1 (y1) X1 + Δg1 (X2 ) (11)
the matrix differential equation of S2. Similarly, from Lemma 2, for
subsystem (6), it follows that v = y1 and S (t ) = S2 . The second part is ϵ̇2 = [A2 (y2 ) − ΓS2−1 C2T C2 − ΛB1 (Z1) C2]ϵ2 − [ΛB (Z1) C1 + KC1T C1]ϵ1
the term associated to the identification of the load torque, which gain + {Φ − ΛB1 (Z1) C2 Λ}ϵ3 + g2 (X1, y2 ) − g2 (Z1, y2 )
is given by ϖΛS3−1 ΛT C2T and it depends on the solution of the differential
+ ΔA2 (y1) X2 + Δg2 (X2 , y1) (12)
equations of S3 and Λ.

Remark 1. The load torque Tl is estimated using Eq. (6) where the ϵ̇3 = −B (Z1) C1 ϵ1 − [ϖS3−1 ΛT C2T + B1 (Z1)] C2 ϵ2 − [ϖS3−1 ΛT C2T + B1 (Z1)] C2
term B (Z1)(y1 − yl1) + B1 (Z1)(y2 − yl2 ) can be expressed, in terms of the Λϵ3. (13)

electromagnetic torque Te and its estimated Te , as follows:
From Lemma 1, there exist t0 ≥ 0 and real numbers ηSmax > 0,
∼ i
B (Z1)(y1 − yl1) + B1 (Z1)(y2 − yl2 ) = k (Te − Te ). ηSmin > 0 , which are independent of ρi, such that V (t , ϵi) = ϵTi Si ϵi for
i
i = 1, 2, 3; satisfies
Furthermore, to estimate the rotor position, the following observer is
designed: ηSmin
i
ϵi 2 ≤ V (t , ϵi) ≤ ηSmax
i
ϵi 2 , ∀ t ≥ t0. (14)

dθl l + Kθ (iq − i^q ) Then, the following result can be established about the observer

dt (7) convergence in spite of bounded parametric uncertainties.

where Kθ is the observer gain. Theorem 1. Consider the IPMSM dynamic model represented by (3)
and (4). System (5) and (6) is an adaptive interconnected observer for
system (3) and (4) with strongly uniformly practical stability of the
5. Observer convergence
estimation error dynamics.
The electrical machine parameters may vary which affects the Proof of Theorem 1. Consider the following Lyapunov function
operation of the machine. The influence of the parameter uncertainties
on the proposed will be studied in this section. The practical stability Vo = ϵ1T S1 ϵ1 + ϵT2 S2 ϵ2 + ϵT3 S3 ϵ3.
theorem introduced in Lakshmikantham, Leela, and Martynyuk (1990)
will be used to ensure the stability of the proposed observer under Then, taking the time derivative and replacing the suitable terms, it
parameter uncertainties (see the Appendix A for more details). From follows that
system (1), the IPMSM parameters are defined in the physical domain V˙o = −ρ1 ϵ1T S1 ϵ1 − ρ2 ϵT2 S2 ϵ2 − ρη ϵT3 S3 ϵ3 + ϵ1T S1 {g1 (X2 , u ) − g1 (Z 2, u )}
+ (Definition 1) with well-known bound values. The interconnected
IPMSM uncertain model is presented as follows: + ϵT2 S2 {[A2 (y1) − A2 (yl1)] X2 + g2 (X1, u , y2 ) − g2 (Z1, u, yl2 )

⎧ X˙1 = A1 (y ) X1 + g (X2 ) + Φ1 (u ) + ΔA1 (y ) X1 + Δg (X2 ) + ΔA2 X2 + Δg2}. (15)


Σ1: ⎨ 1 1 1 1
⎩ y1 = C1 X1 (8) According to Lemma 1 and taking the initial conditions of the IPMSM
drive and of the observer in the physical operation domain + , the
⎧ X˙2 = A2 (y ) X2 + g (X1, y ) + Φ 2 (u ) + ΦTl + ΔA2 (y ) X2 + Δg (X1, y )
Σ 2: ⎨ 1 2 2 1 2 2 following inequalities hold
⎩ y2 = C2 X2
S1 ≤ μ1, S2 ≤ μ2 ,
(9)
S3 ≤ μ3 , g2 (X2 , u , y2 ) − g2 (Z 2, u, yl2 ) ≤ k1 ϵ1 A2 (y1) − A2 (yl1)
where ΔA1 (y1), ΔA2 (y1), Δg1 (X2 ) and Δg2 (X1, y2 ) are respectively the
uncertain terms of A1 (X2 , y ), A2 (X1), g1 (X2 ), g2 (X1, y2 ). The uncertain ≤ k3 ϵ1 g1 (X2 , u ) − g1 (Z 2, u ) ≤ k 4 ϵ2 ΔA2 X2 + Δg2 ≤ k5,
terms are detailed as follows: X2 ≤ k 6. (16)
⎡ ⎛ ⎞ ⎤ ⎡ ⎛ ⎞ ⎛Φ ⎞ ⎤
⎢ 0 − Δ ⎜ 1 ⎟ iq ⎥ ⎢− pΔ ⎜ Ld ⎟ Ωid − pΔ ⎜ f ⎟ Ω ⎥ Now, substituting Eq. (16) into (15), and from Assumption 1, by
ΔA1 (·) = ⎢ ⎝ Lq ⎠ ⎥, Δg1 (·) = ⎢ ⎝ Lq ⎠ ⎝ Lq ⎠ ⎥, regrouping the terms with respect to ϵ1 , ϵ2 and ϵ3 . Then, the time
⎢⎣ 0 0 ⎥⎦ ⎢⎣ 0 ⎥⎦
derivative of Vo (15) satisfies the condition

67
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

V˙o ≤ −ρ1 ϵ1T S1 ϵ1 − ρ2 ϵT2 S2 ϵ2 − ρη ϵT3 Sη ϵ3 + λ1 ϵ1 ϵ2 + λ2 ϵ2 ϵ3 λ max (S )


∥ ϵ(t )∥ ≤ δ* = ϵ*, ∀ t ≥ To
2 λ min (S ) (27)
+ λ3 ϵ2 + λ 4 ϵ2 (17)
and
where λ1 = {μ1 k 4 + μ2 k1 + μ2 k 6 k3}, λ2 = μ2 k2 , λ3 = μ2 k 4 , λ 4 = μ2 k5.
From inequalities (14) and (17), we can be rewritten in terms of V1, 2λ max (S ) ⎛ ∥ ϵ(0)∥ ⎞
To = ln ⎜ ⎟.
V2 and V3 as follows γ ⎝ δ* ⎠
∼ ∼ ∼
V˙o ≤ −ρ1 − V1 − ρ2 V2 − ρη V3 + 2λ1 V1 V2 + 2λ2 V2 V3 + 2λ3 V2 This guarantees that the estimation error converges to the set B (0, ϵ*)
∼ i.e. the ball centered at the origin and radius ϵ*.
+ 2λ 4 V2 (18)
where 6. Robust optimum integral backstepping control of IPM
∼ λ1 ∼ λ2 ∼ λ3 ∼ λ4 synchronous machine
λ1 = , λ2 = , λ3 = min , λ4 = .
ηSmin ηSmin ηSmin ηSmin ηS2 ηSmin
1 2 2 η 2 Several approaches have been proposed for the design of nonlinear
system controllers. The backstepping technique is a systematic and
Furthermore, using the following inequalities
recursive design. The backstepping control principle is based on select
φ1 1 φ2 1 recursively some appropriate function of state variables as a virtual
V1 V2 ≤ V1 + V2, V2 V3 ≤ V2 + V3,
2 2φ1 2 2φ2 control input for a subsystem of order one. The virtual control is
∀ φi ∈ ]0, 1[, (i = 1, 2) (19) designed to obtain the stability of this subsystem via a Lyapunov
function. After each stage, it results a new virtual control obtained from
and by substituting (19) into (18), we obtain the previous step. Backstepping procedure terminates when the real
⎛ input appears. A whole Lyapunov function is defined by summing all
∼⎞
∼ ∼
∼ ∼ λ λ ∼
V˙o ≤ −(ρ1 − λ1 φ1) V1 − ⎜⎜ρ2 − λ2 φ2 − 1 − λ3⎟⎟ V2 − (ρη − 2 ) V3 + λ 4 ϵ2 , previous Lyapunov functions introduced at each step.
⎝ φ1 ⎠ φ2
On the other hand, classical controllers fail to achieve the control
(20) objectives, due to the system constraints (for example: parameter
which can be written as follows uncertainties). In this section, the classical backstepping-based con-
troller is modified by adding integral actions at each step of the
V˙o ≤ −δ (V1 + V2 + V3) + μ ( V1 + V2 ) ≤ −δVo + μψ Vo , (21) algorithm. Then, the modified backstepping controller ensures high
performances by eliminating steady-state error that could be induced
with
otherwise by some disturbances or parametric variations at each level

δ = min(δ1, δ 2, δ3), μ = λ4 of the system structure. The controller consists of three loops.
∼ ∼ ∼
λ1 ∼ ∼
λ2
Speed loop. In the first step, the rotor speed is forced to track its
where δ1 = ρ1 − λ1φ1 > 0 , δ 2 = ρ2 − λ2 φ2 − φ1
− λ3 > 0 , δ3 = ρη − φ2
> 0, reference. The speed tracking error zΩ defined as:
and ψ > 0 , such that t

ψ V1 + V2 + V3 > V1 + V2 .
z Ω = Ω* − Ω + k′Ω ∫0 (Ω* − Ω ) dt
(28)
t
Then, the following inequalities hold with k′Ω ∫ (Ω* − Ω ) dt represents an integral action introduced to
0
guarantee the rotor speed tracking despite parameter uncertainties.
⎧ ρ > λ∼ φ ,
⎪ 1 1 1 In order to ensure the speed tracking, iq will be forced to track iq* by
⎪ ∼ means of the control law vq.
⎪ ρ2 > λ2 φ2 + λ1 + λ∼3,

⎨ φ1 Computing the time derivative of (28), and replacing the current iq
⎪ ∼ by iq*, it follows that:
⎪ λ2
⎪ ρη > φ . p f p 1
⎩ 2 (22) z˙Ω = Ω˙* − (L d − Lq ) id iq* + v Ω − ϕf iq* + Tl + k′Ω (Ω˙* − Ω˙ ).
J J J J (29)
Furthermore, it follows that there exist positive constants, λ min (S ) and
λ max (S ), where S = diag (S1, S2, S3) and ϵ(t ) = (ϵ1T (t ), ϵT2 (t ), ϵT3 (t ))T , such To study the stability of the speed loop, the following candidate
1
that Lyapunov function is chosen as VΩ = 2 zΩ2 . The time derivative of the
above Lyapunov candidate function can be computed as follows:
λ min (S )∥ ϵ(t )∥2 ≤ Vo = ϵ(t )T S ϵ(t ) ≤ λ max (S )∥ ϵ(t )∥2 .
⎧ p f p 1 ⎫
Then, it is obtained V˙Ω = z Ω ⎨Ω˙* − (L d − Lq ) id iq* + v Ω − Φf iq* + Tl + k′Ω (Ω˙* − Ω˙ ) ⎬.
⎩ J J J J ⎭
V˙o ≤ −δVo + μψ Vo , ≤−δλ min (S )∥ ϵ(t )∥2 + μψ λ max (S ) ∥ ϵ(t )∥. (23) (30)
Taking γ such that 0 < γ < δλ min (S ), then the above inequality can be To ensure its negativity and then, the stability of the speed loop, iq* is
expressed as follows chosen as:

J ⎡ fv 1 ⎤
V˙o ≤ −γ ∥ ϵ(t )∥2 + ∥ ϵ(t )∥{(γ − δλ min (S ))∥ ϵ(t )∥ + μψ λ max (S ) }. (24) iq* = ⎢kΩ z Ω + Ω˙* + Ω + k′Ω (Ω* − Ω ) + Tl ⎥ .
p ⎣ J J ⎦
pϕf + (L d − Lq ) id
Then, using this inequality, it is easy to show that J
(31)
V˙o ≤ −γ ∥ ϵ(t )∥2 (25)
Then V˙Ω = −kΩ zΩ2 , with kΩ > 0 . Hence, the speed tracking is ensured.
μψ λmax (S )
if ∥ ϵ(t )∥ ≥ (δλmin (S ) − γ )
= δ*, and from Theorem 5.1 and Corollary 5.3 in Current iq loop. First the following tracking error is defined
[? ], if ∥ ϵ(0)∥ = δ*, there exists To, such that
z q = iq* − iq + z′q (32)
γ
λ max (S ) − t t
∥ ϵ(t )∥ ≤ ∥ ϵ(0)∥ e (2λmax (S ) , ∀ t < To with z′q is an integral action z′q = k′q ∫ (iq* − iq ) dt added to robustify
λ min (S ) (26) 0
the current tracking. The following Lyapunov candidate function is
and chosen to study the stability of the current loop as:

68
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

1 2 1 2 to ensure its negativity. The control law vd which make the d axis
Vq = VΩ + zq + zq ′.
2 2 (33) current loop stable is chosen as follows:
From the mathematical model of the IPMSM, the time derivative of Vq ⎡ t R Lq ⎤
can be computed as follows. To ensure its negativity and then the vd = L d ⎢kd (id* − id + k′d

∫0 (id* − id ) dt ) + id* + s iq − p iq Ω⎥ .
Ld Ld ⎦ (43)
stability of the current loop, the control law vq is chosen as:
The time derivative of Vd is then:
⎡ ϕf R diq* ⎤
vq = Lq ⎢kq z q + p Ω + s iq + ⎥.
V˙d = −(kd − k′d ) z d2 − k′d z d2′.
⎢⎣ Lq Lq dt ⎥⎦ (34)
From (32) and (34) it follows that By choosing Kd = min{(kd − k′d ), k′d }, it follows that
V˙d ≤ −Kd Vd .
V˙q = −kΩ zΩ2 − (kq − k′q ) zq2 − k′q zq2′ ≤ −Kq Vq (35)
Then, the control input vd force the d axis current to track its reference
where Kq = min{kΩ, (kq − k′q ), k′q }. The control law vq ensures that the
computed by MTPA strategy which allows to achieve the second control
current iq tracks the virtual input iq*. This virtual input is computed to
objective.
assure the speed tracking, i.e. (Ω → Ω*). From where, the first control
objective is reached.
Current id loop. 7. Observer-controller scheme stability analysis
The Maximum Torque Per Ampere (MTPA) strategy is proposed to
obtain the most of reluctance torque in the IPMSMs. In the last years, In the previous section, the control laws are computed using a
some effective proposals based on two schemes of the MTPA control perfect system state knowledge. Even if an observer is designed to
have been presented (Uddin & Rahman, 2007). Here, the strategy reconstruct the system state, in practice the observer cannot provides
adopted to obtain the current id reference is based on the mathematical ideal estimates due to disturbances and uncertainties. It is why, in this
model of the IPMSM. The IPMSM electromagnetic torque is given by: section, the stability of the whole system (observer-controller) will be
studied based on the practical stability theory (Lakshmikantham et al.,
Te = p {ϕf iq + (L d − Lq ) id iq}. (36) 1990). In the appendix (Section Appendix A) the necessary definitions
are recalled. To ensure the stability of the observer-controller scheme,
The second term of the electromagnetic torque expression in (36) is
the observation dynamics are taken into account. Following the proof
subject of several researches in order to guarantee an energy efficient
in Hamida et al. (2012), we consider the following candidate Lyapunov
control of the IPMSM. The MTPA strategy (Rahman, Vilathgamuwa,
function:
Uddin, & Tseng, 2003) is developed to provide a maximum torque/
current ratio. The MTPA strategy is now summarized. The stator phase 1 2 1 1 1 1
Voc = Vo + Vc = zΩ + zq2 + zq2′ + z d2 + z d2′ + ϵ1T S1 ϵ1 + ϵT2 S2 ϵ3
Ia current is composed from id and iq as follows: 2 2 2 2 2
Ia2 = iq2 + id2. + ϵT3 S3 ϵ3. (44)
(37)
Taking into account the observation dynamics in the control laws
From (36) and (37) the electromagnetic torque can be expressed in
and by adding ± vq (X ) and ± vd (X ), the time derivative of Vc is obtained
terms of id and Ia:
by:
Te = p [ϕf + (L d − Lq ) id ] Ia2 − id2 . (38) 1
V˙c = −kΩ zΩ2 − {kq − k′q } zq2 − k′q zq2′ − l ) − vq (X )}
z q {vq (X
The derivative of the electromagnetic torque with respect to the current Lq
id is given by: 1 l ) − vd (X )}.
− {kd − k′d } z d2 − k′d z d2′ − z d {vd (X
Ld (45)
∂Te −ϕf id + (L d − Lq )(Ia2 − 2id2 )
=p . Now, considering the following inequalities:
∂id Ia2 − id2 (39)
ξ j1 2 1 ξj2 2 1
From 31, the torque is maximum when
∂Te
= 0 , i.e. |zj | ϵ1 s1 ≤ ϵ1 s1 + |zj |2 |zj | ϵ2 s2 ≤ ϵ2 s2 + |zj |2 |zj
∂id 2 2ξ j1 2 2ξ j 2
ϕf ξj3 1
2id2 + id − Ia2 = 0. | ϵ3 s3 ≤ ϵ3 2
s3 + l ) − vq (X )
|zj |2 |vq (X
(L d − L q ) (40) 2 2ξ j 3

Then, by solving the second order equation (40) the d-axis current | ≤ L1 { ϵ1 s1 + ϵ2 s2 + ϵ3 s3 }|vd (X )
l − vd (X )
reference is given by: | ≤ L 2 { ϵ1 s1 + ϵ2 s2 + ϵ3 s3 } (46)
ϕf ϕf2 ∀ ξ j1, ξ j 2, ξ j 3 ∈ ]0 1[ ; for j=q, d; and by substituting equation (46) into
id* = − − + iq2 .
2(L d − Lq ) 4(L d − Lq )2 (41) (45), we obtain

Now, in order to guarantee an energy efficient operation of the IPMSM V˙oc ≤ −δVo + μφ Vo + ϑ1 ϵ1 2
s1 + ϑ2 ϵ2 2
s2 + ϑ3 ϵ3 2
s3 − ϑ4 zΩ2 − ϑ5zq2
via the MTPA strategy, the current id will be forced to track its − ϑ6 zq2′ − ϑ7z d2 − ϑ8z d2′ (47)
reference given by (41). The d axis current tracking error is defined as:
L ξL1ξq1 L1ξq2 L ξ L1ξq3 L2 ξd 3
where ϑ1 = 2Lq + 22Ldd1 , ϑ2 = 2Lq + 22Ldd 2 , ϑ3 = + , ϑ4 = kΩ ,
z d = id* − id + z′d , 2Lq 2Ld
L1 1 1 1
t ϑ5 = {kq − k′q } − 2Lq { ξ + ξ + ξ }, ϑ6 = k′q , ϑ7 = {kd − k′d }−
q1 q2 q3
the integral term z′d = k′d ∫ (id* − id ) dt is added to eliminate the L2 1 1 1
0 { + ξ + ξ }, 8 ϑ = k ′d.
current tracking error. 2Ld ξ d1 d2 d3

The following candidate Lyapunov function is defined to study the By taking ϑO = max(ϑ1, ϑ2, ϑ3) and ϑC = min(ϑ4, ϑ5, ϑ6, ϑ7, ϑ8), then
stability of the current id loop: the inequality (47) becomes:

1 2 1 2 V˙oc ≤ −ηVoc + μφ Voc (48)


Vd = z d + z d ′.
2 2 (42)
with η = min(δ − ϑO, ϑC).
Its time derivative can be computed and thus a control can be defined Let voc = 2 Voc , then the time derivative of voc is

69
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

v˙oc ≤ −ηvoc + φμ Table 1


Nominal motor parameters.
for which the solution is
Current 6A Torque 9Nm
φμ
νoc (t ) ≤ νoc (t0 ) e−η (t − t 0) + (1 − e−η (t − t 0) ).
η (49) Speed 2100 rpm ϕf 0.18 Wb
Rs 1.2 ohm p 5
The strongly uniformly practical stability of the observer-controller Ld 9.2 mH Lq 5.7 mH
scheme can be proved in the same manner as the observer stability, J 0.0073 kg. m2 fv 0.0034 kg. m2. s−1
that implies the corresponding strongly uniformly practical stability of
the system. Hence, the estimation and tracking errors of the closed-
φμ
loop system converge towards a ball B= oc of radius = oc with = oc = η .
This reasoning can be summarized by the following theorem. 410 V, feeds the PMSM. The PWM frequency is set to 12 kHz. The
pulse-width modulated signals (Agarwal & Agarwal, 2012) were
Theorem 2. Consider the IPMSM dynamic model (1) with the generated on the DS1103 board (Al Nabulsi & Dhaouadi, 2012). The
reference signal Ω*. By using the estimates provided by the results are recorded using the DSP interfaced with the measurement
adaptive interconnected observer (5)–(6) to compute the control equipment and a PC. An incremental encoder was used to obtain the
laws (43) and (31)–(34), then, the closed-loop system, is strongly position and the speed that are solely used for comparison.
uniformly practically stable. The stator resistance of the motor can vary depending on the
operation condition. In order to improve the sensorless control
8. Experimental results performance, the stator resistance is online identified under sensorless
operation, and its value is updated in the observer and the controller
The effectiveness and the performance of the proposed sensorless schemes. The observer and controller parameters are chosen to satisfy
Integral Backstepping control of the IPMSM are tested experimentally the convergence conditions:
in the motor set-up (see Fig. 2) following the designed controller- Controller parameters values:
observer scheme as shown in Fig. 3. Of course, the mechanical sensors
kΩ = 350, k′Ω = 45, kq = 400, k′q = 32, kd = 420, k′d = 65
are only used to validate the estimations.
Observer parameters values:
8.1. PMSM set-up (www2.irccyn.ec nantes.fr/bancessai/)
ρ1 = 400, ρ2 = 500, ρ3 = 120, ϖ = 15, kc1 = .1, kc2 = 0.01,
A DSPACE ADS1103 DSP board was used to carry out the real-time α = 0.01.
algorithm (note that a FPGA also can be used to implement an
The motor is tested according to industrial test trajectories (see
observer-control scheme Alecsa, Cirstea, & Onea, 2012). The para-
www2.irccyn.ec nantes.fr/bancessai/ for details). The trajectories of
meters of the motor are listed in Table 1. The sampling period was set
this benchmark are chosen to test the motor in a wide speed range with
to 100 μs . A three phase IGBT inverter, supplied by a DC link voltage of
the nominal load torque, which are briefly recalled here.
From Fig. 4, at initial time the speed and the load torque values are
zero. Then, the reference speed is carried to 40 rad/s and the load
torque is applied from 2.2 to 3.4 s. This first step allows to test the
performance and the robustness of the observer at low speed. From
4.4–5.4 s, the speed is carried out to 160 rad/s and remains constant
until 7.2 s with the load torque applied from 6.4 to 9.2 s. This second
step is defined to test the observer during a great transient speed and
it's robustness at high speed. Then, the motor is driven to reach zero
speed from 13 s to 16 s. In this time range, the sensorless control law is
tested at zero speed with the applied load torque and without load
torque from 9.2s to 10.4 s. The last step is to test the motor with
nominal load torque in both directions.

8.2. Experimental study


Fig. 2. Motor set-up.
First, the proposed observer-controller scheme has been tested and
implemented using the off line identified parameters of the PMSM
given in Table 1. This case is noted “nominal case” in the following even
if the real parameters are slightly different. Furthermore, the perfor-
mance of the control scheme is tested using industrial test trajectories
of the speed and load torque plotted in Fig. 4a and b, respectively.
The speed estimation performance is shown in Fig. 5, where actual
and estimated speed are shown in Fig. 5a, and the estimation error is
plotted in Fig. 5b. From these Figures, one can see the efficiency of the
proposed adaptive interconnected observer.
The measured (used only for comparison) and estimated positions
are shown in Fig. 6, while the position at very low and zero speed is
plotted in Fig. 7.
The performance of the observer at slow speed is analyzed.
Furthermore, from Fig. 3, in the time interval [8,10] s, the rotor speed
becomes equal to zero. At this time, the observer works, but due to
Fig. 3. Observer-controller scheme. uncertainties and disturbances, the estimation error (see Fig. 5) is

70
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

200

Speed (rad/s)
150
100
50
0
−50
2 4 6 8 10 12 14
Time (s)
10
Torque (rad/s)
5

−5

−10
2 4 6 8 10 12 14
Time (s)

Fig. 4. Industrial benchmark trajectories.

200 10
Ω obs
Ω real 8

150 6

Speed error (rad/s)


4
Speed (rad/s)

100 2

0
50
−2

−4
0
−6

−8
−50
−10
0 5 10 15 0 5 10 15
Time (s) Time (s)
Fig. 5. Speed tracking: nominal case. (a) real and estimated speeds, (b) speed error.

a
0

−1

−2
Position (rad)

−3

−4

−5

−6

−7 θ real θ est
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (s)

Fig. 6. Measured and estimated positions: nominal case.

different to zero, but is bounded. After 10 s, the rotor speed takes a 8.2.1. Robustness study
high value, and then the estimation is improved. This response Now, the robustness of the sensorless control strategy is tested. The
confirms the effectiveness of the proposed observer in this specific torque estimation is shown in Fig. 9a, where one can see the good
operation condition (noted nominal case). estimation when an unknown load torque is applied (see the trajectory
Furthermore, the convergence of the estimates to the actual values of the load torque given in Fig. 4). The error torque observation is given
is ensured and the estimates are bounded. For high rotor speeds the in Fig. 9b (Fig. 10).
convergence of the position estimation to the actual values is verified. It is well-known that the machine parameters can be deviated from
In Fig. 8, the rotor resistance estimation is shown. This estimation the nominal values under temperature variations or magnetic circuit
is used to improve the robustness of the proposed observer-controller saturation. Then, taking into account these parametric deviations in the
scheme. observer-controller scheme, the proposed sensorless control strategy is
experimentally tested with the benchmark reference trajectories

71
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

−1

−2

Position (rad)
−3

−4

−5

−6
θ mes θ obs
−7
8 8.5 9 9.5 10 10.5 11
Temps (s)

Fig. 7. Measured and estimated positions: nominal case, zoom.

1.41

1.405
Resistance (ohm)

1.4

1.395

1.39
0 5 10 15
Time (s)

Fig. 8. Rs estimation: nominal case.

10 2

8
1.5
6
1
Torque error (N.m)

4
Torque (N.m)

0.5
2

0 0

−2
−0.5
−4
−1
−6
Tl −1.5
−8
T l −obs
−10 −2
0 5 10 15 0 5 10 15
Time (s) Time (s)
Fig. 9. Torque estimation: nominal case. (a) real and estimated torques, latex (b) torque error.

(Fig. 4). More precisely, the proposed sensorless control strategy is All these results show clearly the effectiveness of the developed
tested under inductances variations. The system performances under sensorless control strategy.
+20% and −20% variations on the inductance values are respectively
shown on Figs. 11–14. As mentioned before, the inductances affect the 8.3. Comparative study and implementation
behavior to estimate the rotor position when the EMF methods are
implemented to estimate the position. For a more detailed robustness Now, to illustrate the performance of the proposed control scheme,
analysis, see Marwa Ezzat and Glumineau (2010). Then, we can a comparative study will be presented. Here, the Integral Backstepping
conclude that the proposed methodology performs well under these Control proposed in this paper and the High Order Sliding Mode
operating conditions. controller analyzed in Hamida, Glumineau, and de Leon (2013) are
The quality of the load torque estimation is clearly displayed in implemented.
Figs. 9b, 12b, 14b for the nominal and robustness test cases. The It is worth mentioning that the implementation of both control
quality of the tracking is also easy to check in Figs. 5b, 11b, 12b, 13b strategies have been made on the same motor set-up, using the same
also for the nominal and robustness test cases. hardware and software and under the same operation conditions and

72
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

6 400

5
300

4
200

Voltage (V)
Current (A) 3
id
100
iq
2

0
1

0 −100
ud
uq
−1 −200
0 5 10 15 0 5 10 15
Time (s) Time (s)
Fig. 10. Input signals (a) d − q voltages, (b) d − q currents.

200 10
Ω real
Ω obs 8
150 6

Speed error (rad/s)


4
Speed (rad/s)

100 2

0
50 −2

−4
0
−6

−8
−50
−10
0 5 10 15 0 5 10 15
Time (s) Time (s)

Fig. 11. Speed tracking: robustness w.r.t. +20% Ld-Lq deviation. (a) real and estimated speeds, (b) speed error.

10 2

8 1.5
6
1
Torque error (N.m)

4
Torque (N.m)

0.5
2

0 0

−2 −0.5
−4
−1
−6
Tl −1.5
−8
T l −obs
−10 −2
0 5 10 15 0 5 10 15
Time (s) Time (s)

Fig. 12. Torque estimation: robustness w.r.t. +20% Ld-Lq deviation. (a) real and estimated torques, (b) torque error.

the industrial test trajectories. Sliding Mode Controller (see Hamida et al., 2013 for more details) in
From the experimental results obtained, we can remark that. closed-loop with either an Adaptive Interconnected Observer (given in
Hamida et al., 2012) or a HOSM Observer (see Hamida et al., 2014), is
(1) The Integral Backstepping Control shows good transient perfor- presented to illustrate that the proposed method is computationally
mance. satisfactory.
(2) For comparison the integral Backstepping Control is less sensitive The running time of each Algorithm (Controller +Observer) was:
to the noise than switching control as High Order Sliding Mode 38 μs for the Backstepping strategy and 58 μs for Sliding Mode strategy.
Controller even if the high frequency content is attenuated with On the other hand, the running time using PI+FOC strategy (28 μ) is
respect to the classical sliding mode control (Hamida et al., 2013). less than Backstepping strategy.
Notice that the computation time required by the Backstepping
In Table 2, a computation time comparison between a FOC-PI Controller is similar to a classical FOC-PI Controller which is widely
controller, an Integral backstepping Controller and a High Order used in the industrial applications with low-cost hardware. Moreover,

73
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

200 10
Ω obs
Ω real 8
150 6

Speed error (rad/s)


4

Speed (rad/s)
100 2

0
50 −2

−4
0
−6

−8
−50
−10
0 5 10 15 0 5 10 15
Time (s) Time (s)

Fig. 13. Speed tracking: robustness w.r.t. −20% Ld-Lq deviation. (a) real and estimated speeds, (b) speed error.

10 2
8
1.5
6
1
Torque error (N.m)
4
Torque (N.m)

2 0.5

0 0
−2
−0.5
−4
−1
−6

−8 Tl −1.5
T l −obs
−10 −2
0 5 10 15 0 5 10 15
Time (s) Time (s)

Fig. 14. Torque estimation: robustness w.r.t. −20% Ld-Lq deviation. (a) real and estimated torques, (b) torque error.

implemented in real time. Note that the dSPACE implementation


Table 2 allows us to test the feasibility and performance of the proposed
Time computation comparison and mean square error. scheme. However, the algorithms could be implemented in other high-
performance digital signal processor (i.e. Digital Signal Processors or
Controller Control Observation time Mean square error
time (μs) (μs) (rad/s) Field-Programmable Gate Arrays) which are readily available in the
market. On the other hand, the TI TMS320F28335 DSP is one the most
PI 09 19a 0.035 widely used in industrial control applications due to its advanced
Backstepping 10 28b 0.01
properties such as the high speed real-time signal processing with a
HOSM 30 28b 0.047
clock frequency of 150 MHz. The proposed algorithms could be
a
HOSM observer. implemented using this device. The mean square error for each
b
Adaptive interconnected observer. controller is included in Table 2. The backstepping controller offers
the smallest error. This result confirms the high performance of the
the Adaptive Interconnected Observer design is more complex than a proposed controller.
HOSM observer. However, the Adaptive Interconnected Observer is
easier to tune since the parameters are linked to the open loop dynamic 9. Conclusion
of the system.
From the presented data, the PI controller requires the less In this paper, a nonlinear sensorless control for IPMS motor is
computational effort (28 μs in total); the Backstepping scheme in- proposed. An improved backstepping method is developed in order to
creases its computational time, mainly because of the observation ensure the rotor speed and the rotor current to efficiently track the
scheme which is about 30% higher than the observation scheme for the Benchmark trajectories references. For the controller implementation,
PI. Finally, the HOSM scheme presents the higher computational a high gain interconnected observer is presented. The closed loop
needs, mainly due to the increased complexity of the controller, which system stability is ensured using a Lyapunov analysis. Experimental
needs about 3 times the running time of the PI scheme. However, given results are obtained that clearly confirm the effectiveness and the
the used sampling time (100 μs) any of the proposed schemes can be performance of the proposed method at different speed ranges under
significant robustness tests.

74
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

Appendix A

This section is devoted to introduce some concepts and results of practical stability properties using in terms of Lyapunov functions
(Lakshmikantham et al., 1990).
Theory of stability is the basis of the control systems study. Moreover, the concept of practical stability allows to study the properties of a
nonlinear system when the state of this system is bringing close to a set instead of the equilibrium point. This is clear that, in practice, asymptotic
stability towards a domain whose the size has to be determined, for performance checking, is sufficient. If the behavior of the system can be bounded
by certain bounds, the notion of practical stability becomes useful.
Now, we introduce definitions which are useful for to guarantee the practical stability, in terms of Lyapunov functions (For more details see
Lakshmikantham et al., 1990).
Define the following class of function
W = {d1 ∈ C [R+, R+]: d1 (l ) is strictly increasing in l and d1 (l ) → ∞ as l → ∞}.
Let Br = {e ∈ Rn: e ≤ r}. Consider the dynamical system
e˙ = f (t , e), e (t0 ) = e0 , t0 ≥ 0, (50)
A result of the practical stability in terms of Lyapunov-like functions can be given from Lakshmikantham et al. (1990). Assume that

(i) = 1 and = 2 are given such that 0 < = 1 < = 2 ,


(ii) V ∈ C [R+ × Rn, R+] and V (t , e) is locally Lipschitz in e,
(iii) for (t , e) ∈ R+ × B= 2 , d1 ( e ) ≤ V (t , e) ≤ d2 ( e ) and
V˙ (t , e) ≤ ℘(t , V (t , e)) (51)
where d1, d2 ∈ W and ℘ ∈ C [R+ × R2, R],
(iv) d2 (= 1) < d1 (= 2 ) holds.

Consequently, the practical stability properties of


V˙ (t , e) ≤ ℘(t , V (t , e)) = −α1 V (t , e) + α2, (52)

with α1 and α2 > 0, V0≔V (0, e (0)) ≥ 0 implies strong uniform practical stability of system (50).
The solution of equation (52) is of the form
α2
V (t , e) ≤ V0 e−α1t − t 0 + [1 − e−α1 (t − t 0) ], t ≥ t0.
α1 (53)

References Institute, 349(5), 1734–1757. http://dx.doi.org/10.1016/j.jfranklin.2012.02.005.


Hamida, M. A., Leon, J. D., & Glumineau, A. (2014). High order sliding-mode observer
and integral backstepping control for sensorless ipms motor. International Journal
Accetta, A., Cirrincione, M., Pucci, M., & Vitale, G. (2012). Sensorless control of pmsm of Control, 87(10), 2176–2193.
fractional horsepower drives by signal injection and neural adaptive-band filtering. Kim, S., Ha, J.-I., & Sul, S.-K. (2012). Pwm switching frequency signal injection
IEEE Transactions on Industrial Electronics, 59(3), sensorless method in ipmsm. Industry Applications, IEEE Transactions on, 48(5),
1355–1366. http://dx.doi.org/10.1109/TIE.2011.2167729. 1576–1587. http://dx.doi.org/10.1109/TIA.2012.2210175.
Agarwal, A., & Agarwal, V. (2012). Fpga realization of trapezoidal pwm for generalized Kshirsagar, P., Burgos, R., Jang, J., Lidozzi, A., Wang, F., Boroyevich, D., & Sul, S.-K.
frequency converter. IEEE Transactions on Industrial Informatics, 8(3), 10. http:// (2012). Implementation and sensorless vector-control design and tuning strategy for
dx.doi.org/10.1109/TII.2012.2193406. smpm machines in fan-type applications. Industry Applications, IEEE Transactions
Alecsa, B., Cirstea, M., & Onea, A. (2012). Simulink modeling and design of an efficient on, 48(6), 2402–2413. http://dx.doi.org/10.1109/TIA.2012.2227135.
hardware-constrained fpga-based pmsm speed controller. IEEE Transactions on Lakshmikantham, V., Leela, S., & Martynyuk, A. (1990). Practical stability of nonlinear
Industrial Informatics, 8(3), 9. http://dx.doi.org/10.1109/TII.2012.2193891. systems. World Scientific Publishing.
Alwi, H., & Edwards, C. (2014). Second order sliding mode observers for the {ADDSAFE} Liu, J., & Zhu, Z. (2014). Novel sensorless control strategy with injection of high-
actuator benchmark problem. Control Engineering Practice, 31, frequency pulsating carrier signal into stationary reference frame. IEEE
74–91. http://dx.doi.org/10.1016/j.conengprac.2013.09.014. Transactions on Industry Applications, 50(4),
Besancon, G., & Hammouri, H. (1996). Observer synthesis for class of nonlinear control 2574–2583. http://dx.doi.org/10.1109/TIA.2013.2293000.
systems. European Journal of Control, 2, 176–192. Liu, J., & Zhu, Z. (2014). Improved sensorless control of permanent-magnet synchronous
Besancon, G., & Hammouri, H. (1998). On observer design for interconnected systems. machine based on third-harmonic back emf. IEEE Transactions on Industry
Journal of Mathematical Systems Estimation and Control, 8, 1–25. Applications, 50(3), 1861–1870. http://dx.doi.org/10.1109/TIA.2013.2284299.
Besancon, G., De Leon, J., & Huerta, O. (2006). On adaptive observers for state affine Liu, J., Laghrouche, S., Harmouche, M., & Wack, M. (2014). Adaptive-gain second-order
systems. International Journal of Control, 79, 581–591. sliding mode observer design for switching power converters. Control Engineering
Betin, F., Capolino, G.-A., Casadei, D., Kawkabani, B., Bojoi, R., Harnefors, L. Fahimi, B. Practice, 30, 124–131. http://dx.doi.org/10.1016/j.conengprac.2013.10.012.
(2014). Trends in electrical machines control: Samples for classical, sensorless, and Lu, K., Lei, X., & Blaabjerg, F. (2013). Artificial inductance concept to compensate
fault-tolerant techniques. IEEE Industrial Electronics Magazine, 8(2), nonlinear inductance effects in the back emf-based sensorless control method for
43–55. http://dx.doi.org/10.1109/MIE.2014.2313752. pmsm. Energy Conversion, IEEE Transactions on, 28(3),
Delpoux, R., Bodson, M., & Floquet, T. (2014). Parameter estimation of permanent 593–600. http://dx.doi.org/10.1109/TEC.2013.2261995.
magnet stepper motors without mechanical sensors. Control Engineering Practice, Lu, K., Lei, X., & Blaabjerg, F. (2013). Artificial inductance concept to compensate
26, 178–187. http://dx.doi.org/10.1016/j.conengprac.2014.01.015. nonlinear inductance effects in the back emf-based sensorless control method for
Escareno, J.-A., Rakotondrabe, M., & Habineza, D. (2015). Backstepping-based robust- pmsm. Energy Conversion, IEEE Transactions on, 28(3),
adaptive control of a nonlinear 2-dof piezoactuator. Control Engineering Practice, 593–600. http://dx.doi.org/10.1109/TEC.2013.2261995.
41, 57–71. http://dx.doi.org/10.1016/j.conengprac.2015.04.007. Marwa Ezzat, R. B., & Glumineau, Alain (2010). Comparaison de deux observateurs non
Hamida, M. A., Glumineau, A., & de Leon, J. (2013). High order sliding mode controller linéaires pour la commande sans capteur de la msap: validation expérimentale. In
and observer for sensorless ipm sycnronous motor. In Proceedings of the 4th IEEE Proceedings of Conférence Internationale Francophone d'Automatique. CIFA 2010,
POWERENG conference pp. 7107 –7112. Istanbul, Turky 13–17 May 2013. Nancy, France.
Hamida, M. A., Glumineau, A., & de Leon, J. (2012). Robust integral backstepping Al Nabulsi, A., & Dhaouadi, R. (2012). Efficiency optimization of a dsp-based standalone
control for sensorless {IPM} synchronous motor controller. Journal of the Franklin pv system using fuzzy logic and dual-mppt contro. IEEE Transactions on Industrial

75
M.A. Hamida et al. Control Engineering Practice 59 (2017) 64–76

Informatic, 8(3), 12. http://dx.doi.org/10.1109/TII.2012.2192282. observer. ISA Transactions, 54,


Niapour, S. M., Tabarraie, M., & Feyzi, M. (2014). A new robust speed-sensorless control 15–26. http://dx.doi.org/10.1016/j.isatra.2014.08.008.
strategy for high-performance brushless {DC} motor drives with reduced torque Uddin, M. N., & Rahman, M. A. (2007). High-speed control of ipmsm drives using
ripple. Control Engineering Practice, 24, improved fuzzy logic algorithms. IEEE Transactions on Industrial Electronics, 54,
42–54. http://dx.doi.org/10.1016/j.conengprac.2013.11.014. 190–199.
Pillay, P., & Krishnan, R. (1989). Modeling, simulation, and analysis of permanent- Wang, G., Li, Z., Zhang, G., Yu, Y., & Xu, D. (2013). Quadrature pll-based high-order
magnet motor drives. ii. the brushless dc motor drive. IEEE Transactions on sliding-mode observer for ipmsm sensorless control with online mtpa control
Industry Applications, 25(2), 274–279. http://dx.doi.org/10.1109/28.25542. strategy. Energy Conversion, IEEE Transactions on, 28(1),
Qiao, Z., Shi, T., Wang, Y., Yan, Y., Xia, C., & He, X. (2013). New sliding-mode observer 214–224. http://dx.doi.org/10.1109/TEC.2012.2228484.
for position sensorless control of permanent-magnet synchronous motor. IEEE Wang, G., Zhan, H., Zhang, G., Gui, X., & Xu, D. (2014). Adaptive compensation method
Transactions on Industrial Electronics, 60(2), of position estimation harmonic error for emf-based observer in sensorless ipmsm
710–719. http://dx.doi.org/10.1109/TIE.2012.2206359. drives. Power Electronics, IEEE Transactions on, 29(6),
Raggl, K., Warberger, B., Nussbaumer, T., Burger, S., & Kolar, J. (2009). Robust angle- 3055–3064. http://dx.doi.org/10.1109/TPEL.2013.2276613.
sensorless control of a pmsm bearingless pump. IEEE Transactions on Industrial www2.irccyn.ec/nantes.fr/bancessai/.
Electronics, 56(6), 2076–2085. http://dx.doi.org/10.1109/TIE.2009.2014745. Xu, D., Zhang, S., & Liu, J. (2013). Very-low speed control of {PMSM} based on {EKF}
Rahman, M., Vilathgamuwa, D., Uddin, M., & Tseng, K.-J. (2003). Nonlinear control of estimation with closed loop optimized parameters. ISA Transactions, 52(6),
interior permanent-magnet synchronous motor. IEEE Transactions on Industry 835–843. http://dx.doi.org/10.1016/j.isatra.2013.06.008.
Applications, 39, 408–416. Zhang, Q. (2002). Adaptive observer for multiple-input-multiple-output (mimo) linear
Ren, J.-J., Liu, Y.-C., Wang, N., & Liu, S.-Y. (2015). Sensorless control of ship propulsion time-varying systems. IEEE Transactions on Automatic Control, 47(3),
interior permanent magnet synchronous motor based on a new sliding mode 525–529. http://dx.doi.org/10.1109/9.989154.

76

You might also like