Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

1916 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO.

3, MARCH 2017

Finite-Control Set Model Predictive Control


Method for Torque Control of Induction Motors
Using a State Tracking Cost Index
Abdelsalam A. Ahmed, Member, IEEE, Byung Kwon Koh, Hyo Sung Park,
Kyo-Beum Lee, Senior Member, IEEE, and Young Il Lee, Senior Member, IEEE

Abstract—This paper presents a novel torque control [1]–[5] and for current control of IM and permanent magnet
method for two-level-inverter-fed induction motor drives. synchronous motors (PMSM) [5]–[8]. This approach is called
The control principle is based on a finite-control set model finite-control set MPC (FCS-MPC) and is similar to conven-
predictive control (FCS-MPC) using a state tracking cost
index. In the online procedure of the proposed FCS-MPC, tional direct torque control (DTC) [9] in that an output voltage
the optimal voltage vector and its corresponding optimal vector is applied to the motor during the whole sampling period.
modulation factor are determined based on the principle of The drawback of the conventional MPC is that the time duration
torque and rotor flux error minimization. In this method, a of voltage vector is fixed, which limits the choice of the voltage
reference state is determined in a systematic way so that the
vector in the inverter and produces large ripples in torque and
reference torque tracking with maximum torque per ampere
and flux-limited operation could be achieved. In addition, a current. In order to avoid this drawback of FCS-MPC, numer-
weighting matrix for the state tracking error is optimized in ous research activities have been conducted [10]–[15]. In [10],
offline using the linear matrix inequality based optimization the optimal voltage vector is first determined and then the opti-
problem. The efficacy of the proposed FCS-MPC method is mal duration is calculated according to the principle of torque
proved by the simulation and experimental results at dif-
ripples minimization. However, it presents poor low-speed per-
ferent working circumstances. The comparison of the pre-
sented control system with the conventional FCS-MPC and formance and high-current ripples. FCS-MPC in [12] computes
with other reported FCS-MPC with modulation control is optimal modulation factors for all nonzero vectors following
made. The proposed algorithm yields fast dynamic perfor- deadbeat criteria, and then obtains the optimal control vector
mance and minimum torque and current ripples at different with the optimal modulation factor considering a cost index
speed and torque levels.
penalizing torque and flux tracking error separately. In [15], a
Index Terms—Finite-control set model predictive control simplified FCS-MPC based on a new predictive torque control
(FCS-MPC), flux-increased and flux-limited control, induc- switching table was proposed to reduce the numbers of voltage
tion motors (IM), linear matrix inequality (LMI), maximum vectors. In this algorithm, only three voltage vectors for predic-
torque/ampere control, modulation factor, torque control.
tion and actuation are employed. The maximum reduction of the
I. INTRODUCTION average execution time and the average switching frequency of
each semiconductor switch are considerable. However, the per-
ECENTLY, model predictive control (MPC) methods have
R been applied for torque control of induction motors (IMs)
formance of [15] is comparable to conventional FCS-predictive
torque control (PTC) without duration control, which suffers
from high ripples in torque and current.
Manuscript received February 22, 2016; revised May 20, 2016 and
June 28, 2016; accepted July 25, 2016. Date of publication November In order to have a low absolute current, a maximum torque per
22, 2016; date of current version February 9, 2017. This work was ampere (MTPA) criteria is adopted in many research activities of
supported in part by the National Research Foundation of Korea (NRF) IM drives. MPC considering MTPA for PMSM were proposed in
under Grant NRF -2015R1D1A1A01060451 and in part by Human Re-
sources Development of the Korea Institute of Energy Technology Eval- [14] and [16]. However, insufficient research activities have been
uation and Planning, Korea Government Ministry of Trade, Industry and conducted on an FCS-PMC method applying MTPA and flux-
Energy, under Grant 20154030200720. (Corresponding author: Byung limited operation for IM. The literature reports different MTPA
Kwon Koh.)
A. A. Ahmed is with the Department of Electrical Power and Mach- control schemes for IM drives [17]–[21]. In these references, a
ines Engineering, Faculty of Engineering, Tanta University, Tanta 31527, complex computation to obtain reference slip speed is required
Egypt (e-mail: abdelsalam.abdelsalam@f-eng.tanta.edu.eg). [17], the performance of the controller is tested only at high
B. K. Koh, H. S. Park, and Y. I. Lee are with the Department of Elec-
trical and Information Engineering Seoul National University of Science speed without load [18].
and Technology , Seoul 01811, South Korea (e-mail: 09122303@seoul- In this paper, a new FCS-MPC method is proposed for torque
tech.ac.kr; soulyuki@seoultech.ac.kr; yilee@seoultech.ac.kr). control of an IM. The novelty of the proposed approach is in
K.-B. Lee is with the Department of Electrical and Computer Engi-
neering, Ajou University, Suwon 16499, South Korea (e-mail: kyl@ajou. the computation of reference state considering the MTPA in
ac.kr). flux-increased and flux-limited operational region and the on-
Color versions of one or more of the figures in this paper are available line transition between the two regions at different torque and
online at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TIE.2016.2631456 speed levels. The reference state consists of the command stator
0278-0046 © 2016 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications standards/publications/rights/index.html for more information.
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1917

current and command rotor flux, which is adapted according to variables can be expressed in state-space form as
the given torque requirement. The predicted state tracking error
ẋ (t) = Ac (ωe (t)) x (t) + Bc u (t) . (3)
for this reference state is used in the cost index. This approach
was not considered in the reported references [1]–[16]. The most The matrices Ac and Bc are state-space matrices, given by
effective elements in this control method are computation of ref- ⎡ 1  ⎤
erence states, weighting factors of the objective index, and the − α Rs + β 2 Rr + jωe β Rr β
α L r + j α ωr
duration interval of the optimized voltage vector. The reference Ac (ωe ) = ⎣ ⎦
states are determined according to the operational modes. βRr − L r + j (ωe − ωr )
Rr

The weighting factors used in the cost index are determined  T


1
systematically using linear matrix inequalities (LMI) based op- Bc = 0 (4)
timization technique unlike the reported scheme [12], where α
the weighting factor was determined in a heuristic way. The where x(t) = [ ids iqs λdr λqr ]T are the internal states, u(t) =
optimal voltage vector minimizing the cost function is chosen [ vds vqs ]T are the inputs, α = Ls − L2m /Lr is the stator tran-
first. Then, the optimal duration of that vector is determined sient inductance, β = Lm /Lr is the rotor coupling factor. It is
according to the principle of torque and rotor flux error min- worth noting that the matrix Ac depends on the instantaneous
imization while the duration interval is optimized depending value of mechanical speed ωr . Hence, a discrete time-varying
only on the tracking of torque in [12]. Delay compensator, state model that can be updated at every sampling interval with the
change minimizer for the reduction of inverter commutation, new measured value of ωr is obtained. In the later derivation,
and parameters uncertainty are adopted. Ac will be used instead of Ac (ωe ) for notational simplicity. The
The control is quite different from the usual model predic- following discrete state–space model of the IM can be obtained
tive torque control because it depends on the concept of field- from (3) using the Euler method with the sampling period h
oriented control (FOC) in addition to the selection of the optimal
voltage vector (similar to DTC). Also, in the presented approach, x [k + 1] = Ax [k] + Bu [k] (5)
voltage and current constraints are guaranteed with feasibility where A = I4×4 + Ac h is a 4 × 4 matrix, B = Bc h is a 2 × 1
study of the system’s states. The proposed MPC for torque con- matrix, x[k] = [ ids iqs λdr λqr ]T are the internal states, u[k] =
trol is validated by comparing the simulation and experimental
[ vds vqs ]T are the inputs. Then, the electromagnetic torque of
results with those of the conventional MPC with full sampling
the IM in terms of the stator current, rotor flux, and number of
interval as in [1]–[8] and with the reported work in [12] with
poles P is given as
modulation control.
3 P Lm
Te = (λdr iqs − λqr ids ) . (6)
II. MODEL OF THE ELECTRICAL SYSTEM 2 2 Lr
A. Dynamic Equations of IM B. Steady-State Control Input
The stator and rotor voltage equations of an IM in the syn- The feasibility of a specific rotational speed for a given ref-
chronous reference frame are as follows [22]: erence state x∗ = [ i∗T ∗T T
s λr ] can be checked by considering a
d steady state of the model (5) i.e.,
vdqs = Rs idqs + λdqs + jωe λdqs (1)
dt x∗ = Ax∗ + Bu∗ (7)
d  ∗     
0 = Rr idqr + λdqr + j (ωe − ωr ) λdqr (2) is A11 A12 i∗s B1 ∗
dt ∗ = ∗ + u (8)
λr A21 A22 λr 0
where
  where
0 1
 
j= 1− Ts
Rs + β 2 Rr ωe Ts
−1 0 α
A11 =  
Rs , Rr , Ls , Lr , Lm , ωe , and ωr denote stator resistance, rotor re- −ωe Ts 1− Ts
α Rs + β 2 Rr
sistance, stator inductance, rotor inductance, mutual inductance, ⎡ ⎤

β Rr β
Ts α ωr Ts βRr Ts 0
= ⎣ ⎦,
and rotational speeds of the stator current and rotor, respectively. αLr
A12 A21 =
Stator and rotor flux linkage can be defined as − αβ ωr Ts β Rr
αLr Ts 0 βRr Ts
λdqs = Ls idqs + Lm idqr

1− Rr
Lr Ts (ωe − ωr ) Ts
λdqr = Lr idqr + Lm idqs . A22 = .
− (ωe − ωr ) Ts 1− Rr
Lr Ts
Given the equations presented above, it is possible to represent
The first row block of (8) can be rewritten as i∗s = A11 i∗s +
the motor behavior using four internal variables (stator currents
A12 λ∗r + B1 u∗ ; therefore, the steady-state control input can be
and rotor fluxes) and two input (stator voltage vectors) [23]. The
given as
dynamic equations of IM (1) and (2), with rotor flux linkage
components λdqr and stator current components idqs as state u∗ = B1−1 {(I2X2 − A11 ) i∗s − A12 λ∗r } . (9)
1918 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

TABLE I With the aid of (6), considering the principle of indirect FOC,
VOLTAGE VECTORS OF A TWO-LEVEL INVERTER
the reference value for the q-axis stator current can be obtained
from the reference torque and flux as
State Sa Sb Sc Voltage vector
2 2 Lr Te∗ [k]
0 OFF OFF OFF v 0 = 0∠0 i∗qs [k] = . (11)
1 ON OFF OFF v 1 = 2/3V d c ∠0 3 P Lm λ∗
2 ON ON OFF v 2 = 2/3V d c ∠60
3 OFF ON OFF v 3 = 2/3V d c ∠120
From (10) and (11), the reference state
4 OFF ON ON v 4 = 2/3V d c ∠180 T
5 OFF OFF ON v 5 = 2/3V d c ∠240 x∗ (Te∗ , λ∗ ) = i∗ds i∗qs λ∗dr λ∗qr
6 ON OFF ON v 6 = 2/3V d c ∠300
7 ON ON ON v 7 = 0∠0 can be defined as follows:
 ∗ T
λ 2 2 Lr Te∗ ∗
x∗ (Te∗ , λ∗ ) = λ 0 . (12)
Lm 3 P Lm λ∗

B. Reference State With Flux-Increased and


Flux-Limited Operation
From the previous theoretical analysis, it can be noticed that
the reference state of (12) depends on the given reference torque
Te∗ and the reference rotor flux λ∗ . The choice of reference flux
Fig. 1. Configuration of the proposed FCS-MPC control of IM drive.
λ∗ will be made to produce the maximum available torque per
ampere as follows.
Note that the control input voltage u∗ depends on the ro- The electromagnetic torque in FOC can be expressed as Te =
tational speed ωr since matrices A11 and A12 are functions 2
k ids iqs , where k = 32 P2 LLmr is the torque constant. To maximize
of ωr . Thus, the system (5) is a nonlinear system. The fea-
the efficiency of the system, both copper loss in rotor and stator
sibility of a rotational speed ωr , for a given reference state
windings and iron loss need to be considered [25]. However,
x∗ , would be determined by the feasibility of the u∗ with the
this is out of scope of this paper. Then, it is easy to see that
available output voltage vectors. Considering the output volt- 1
the condition to minimize |is | = (i2ds + i2qs ) 2 for a given Te ,
age vectors as in Table I [24], the criteria for the feasibility of
without ohmic and iron losses, is that the current components
u∗ = [ u∗d u∗q ]T can be determined in terms of dc-linked voltage
are equal, i.e., ids = iqs [17], [26]. Thus, the MTPA control
for hexagonal and circular boundaries as u∗2 ∗2
d + ud ≤ 3 Vdc
2
could be achieved if the flux reference λ∗ is determined so that

∗2 ∗2
and ud + ud ≤ √3 Vdc , respectively.
1

i∗ds = i∗qs . (13)


III. DESIGN OF PROPOSED FCS-MPC CONTROLLER
From (10) and (11), we can see that the flux reference λ̂∗ (k)
The schematic of proposed MPC is shown in Fig. 1. In the for MTPA operation is
following sections, each block is described in detail. 
∗ 22
λ̂ (k) = Lr Te∗ (k). (14)
A. Definition of Computation of Reference State 3P
The reference torque Te∗ (k) is assumed to be given at each The flux reference in (14) is modified to add a design param-
time step for the torque control of an IM. A method to provide eter η for flexibility of control algorithm
a proper reference state x∗ (Te∗ (k)) for a given reference torque λ∗ (k) = η.λ̂∗ (k) (15)
is proposed. Note that the state x of the system (4) consists of
stator current idqs and rotor flux λdqr . The rotor flux rotates at where η(≤ 1) is a positive design parameter, called here as
the synchronized speed ωe and the rotational angle θe changes torque per ampere ratio. In the constant power operation of the
accordingly. IM under FOC, the rotor flux reference is determined according
The rotating dq-frame used in (1) and (2) is based on the to flux value. In this region, the flux level split in two stages:
rotor flux, i.e., the position of the rotor flux serves as the d- flux-increased and flux-limited subregions. Calculation of η in
axis. Thus, considering the principle of indirect FOC, the direct these two cases is defined as per the following remark.
axis is oriented to the reference rotor flux value, i.e., λ∗dr = λ∗ Remark 1: As stated in the literature, the slip speed of an IM
and the quadrature component of rotor flux becomes zero, i.e., can be described as
λ∗qr = 0. The choice of λ∗ will be considered later in this section 1 iqs
concerning the torque per ampere ratio. In a steady state, the ωsl = (16)
τr ids
stator current is related to the rotor flux, i.e., λdr = Lm ids . Then,
the reference value for the d-axis stator current is determined as where τr is the rotor time constant. Then, using relations (10)
follows: and (11), the optimal constant slip speed can be written as
λ∗ [k] 1 22 T∗
i∗ds [k] = . (10) ωsl∗ = Lr ∗e2 . (17)
Lm τr 3 P λ
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1919

From (17), we have for the proposed control system is chosen as



1 22 J (k) = (e [k + 1])T W e [k + 1] (22)
λ∗ = ∗
Lr Te∗ . (18)
τr ωsp 3P where W = W T is a positive definite weighting matrix.
Comparing (15) and (18), we can see that the torque per
ampere ratio can be set as D. Calculation of Weighting Matrix
1 The choice of the weighting matrix W is important to obtain a
η= ∗
. (19)
τr ωsp good state tracking performance. Here, we propose a method for
choosing W so that J(k) becomes a monotonically decreasing
As observed above, the MTPA operation is obtained when Lyapunov function for the error dynamics of the system (5)
the slip speed is set as η = 1, i.e., when the steady-state input for the reference state is applied,
∗ 1 i.e., u[k] = u∗ . Subtracting (5) from (7) with u[k] = u∗ yields
ωsp = . (20)
τr e[k + 1] = Ae[k]. Thus, the difference between two adjacent
In the flux-limited region [17], however, it is required to values of the cost index can be expressed as follows
increase the slip speed (i.e., decrease η) as the torque reference J (k) − J (k − 1) = eT [k + 1] W e [k + 1] − eT [k] W e [k]
Te∗ increases. Otherwise, the output torque cannot be maintained  
at the desired value because of the limitation on the voltage = eT [k] AT W A − W e [k] . (23)
vector. According to the desired slip speed presented in [17], in From observation of (23), it can be seen that
the flux-limited region η can be chosen as
 W − AT W A > 0 (24)
  

η = 2 Te L Lr / 1 − 1 − 4T e C L2m
2 ∗ 2 (21) guarantees that J(k) − J(k − 1) < 0, i.e., J(k) is monotoni-
cally decreasing when the steady-state input for the reference
where state is applied. Thus, we could consider the cost index J(k) as
Lm Llr a Lyapunov function because it has the properties of Lyapunov
L = Lls + function, i.e., it is positive and monotonically decreasing [29],
Lm + Llr
 [30].
L2 L4m + L2 L2r + 2L2m LLr Multiplying both sides of (24) with Q(= W −1 ) produces
C= .
L4m
Q − QAT Q−1 AQ > 0. (25)

Substituting η of (21) into (19) gives = ωsp .η 2 τ1r
A feasible solution was found for this problem so that stability
It will be shown later that use of (17) and (19) makes η
is guaranteed. After applying the Schur complement, (25) is
decrease as the torque reference Te∗ increases. Thus, we propose
equivalent to the following LMI [22], [23]:
the following criteria to choose η as (20) or (21).

Operational-Mode Selection (OMS) Algorithm: Q QAT
> 0. (26)
Step 1: For a given torque reference, compute the reference state AQ Q
x∗ (k) as (12) and (15) with η = 1. Note that the solution of (26) is not unique and we need to
Step 2: Calculate the input control voltage vectors u satisfying define an optimization problem as follows:
(9) for x∗ (k) of step 1.
Step 3: If the u∗ (k) of step 2 satisfies the feasibility condition min f (Q) subject to (26) (27)
Q
of input voltage, choose η as (20) or determine η as (21).
where f (Q) is a function of Q determined so that (27) can be
Using the above OMS algorithm, the boundary torque be- solved efficiently with the MATLAB LMI toolbox. Thus, for
tween the MTPA and the limited-flux operational modes can torque control on the IM, the weighing factors are optimized of-
be found at a given operational speed. Then, a look-up table fline using the LMI-based optimization problem at each section
containing the boundary torque versus operational speed can be of rotational speed. Then, these values are used in the online im-
established. plementation. The stability of a receding horizon control (RHC)
minimizing the cost index J(k) of (22) at each time step can be
C. Definition of Cost Function established as per the following theorem.
This section presents the cost function of the proposed FCS- Theorem 1: Consider the discrete-time system (5) and a con-
MPC method for the discrete-time system (5). For this purpose, stant reference state x∗ and its corresponding steady-state con-
the one-step ahead predicted error state is defined based on the trol input u∗ satisfying (7). Then, the use of control input u[k]
reference state x∗ (Te∗ , λ∗ ) as minimizing the cost index J(k) of (22) guarantees that the error
state e[k] goes to zero as k increases to infinity provided that the
e [k + 1] = x∗ (Te∗ , λ∗ ) − x [k + 1|k]
weighting matrix W is chosen following (27).
where x[k + 1|k] is the one-step ahead prediction of the state Proof: The argument between (23)–(27) verifies that the use
x[k + 1] obtained from the dynamics (5). Now, the cost function of u∗ will guarantee J(k) − J(k − 1) < 0. Then, it is easy to
1920 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

see that the use of the optimal control input u∗ [k] minimizing Note that (31) is a quadratic equation of modulation factor
J(k) also guarantees J(k) − J(k − 1) < 0. Thus, the use of the μ. Thus, the optimal value of μ that minimizes (31) can be
optimal control input at each time step yields obtained from δδμ Jsel (k) = 0 as (33), shown at the bottom of the
page.
. . . < J (k + 1) < J (k) < J (k − 1) From (33), it can be found that the optimal modulation factor
which guarantees e[k] → 0.  μ∗ is got at time k considering both of torque- and flux-state
The difference between the RHC and the proposed FCS-MPC error. Also, it is noted that the rotor flux {x[k]}3,4 of (5) is not
is that the optimal control input of the RHC can be implemented measurable. Thus, the rotor flux needs to be estimated using
exactly while the control input of the proposed FCS-MPC is measurable signals to compute μ∗ of (33). In order to estimate
chosen among the available six active vectors of the inverter the rotor flux, the following current model of the IM is adopted:
with duration control as will be explained in the next section.  
d Lm Rr
λdqr = Rr idqs − − jωr λdqr . (34)
dt Lr Lr
E. Design of Proposed FCS-MPC With Optimal Duration
In real applications, however, we need to modify this proce-
The control strategy of the proposed MPC comprises two dure as described in the following section to compensate for the
steps: find the optimal output voltage vector vsel from Table I input time delay and steady-state error.
that yields the smallest cost index J(k) of (22) and compute
how long that output voltage vector vsel needs to be applied F. Compensation for Time Delay
during the sampling period h to minimize the cost index. In
order to determine the optimal output voltage vector vsel , the The control voltage u(k) = μ∗ vsel of Section III-E is sup-
control input u[k] of (5) should be substituted with each active posed to be applied to the inverter at time step k to minimize
output voltage vector listed in Table I to produce the predicted the cost index J(k) concerning the state tracking errors at time
value as follow. step (k + 1). In real applications, however, there is a time delay
The system described in (5) is rewritten for all possible volt- due to the computation time and modulation mechanism, i.e.,
age set as the control input computed at time step k is actually applied to
the inverter at time step (k + 1). This time delay may degrade
xi [k + 1|k] = A x [k] + Bvi , i = 1, . . . ., 6. (28) the performance, and we need to compensate for it. The method
for compensating the time delay presented in [31] is used. From
Inserting (28) into the cost index of (22) yields
(5), the prediction x[k + 2|k] can be made as
Ji (k) = (ei [k + 1|k])T W (ei [k + 1|k]) (29)
x [k + 2|k] = Ax [k + 1|k] + Bu [k] . (35)

where ei [k + 1|k] = x (Te∗ , ∗
λ ) − xi [k + 1|k] is the state error Based on (35), the cost index (22) is modified as follows:
for i = 1, . . . ., 6. The output voltage vector vi∗ that yields
the smallest value of Ji (k) for i = 1, . . . ., 6 is chosen to J (k + 2) = (x∗ (Te∗ , λ∗ ) − x [k + 2|k])T W (x∗ (Te∗ , λ∗ )
be vsel (k). Applying vsel over the whole sampling period h
− x [k + 2|k]) . (36)
will produce a high torque ripple, the application time of vsel
is needed to be adjusted. The modulation factor μ is used to Now, the optimal voltage vector vsel should be selected among
adjust the application time of vsel . The predicted state with this the eight voltage vectors to minimize J(k) of (36). In addition,
modulation factor and vsel is given by the modulation factor μ∗ is also modified accordingly to yield
(37), shown at the bottom of the page.
xsel [k + 1|k] = A x [k] + B μ vsel [k] . (30)
Then, the cost index in (29) is augmented as follows: G. Use of a Reference State Integrator

Jsel (k) = esel [k + 1|k]T W esel [k + 1|k] (31) The reference state derived in (8) may not be correct be-
cause of uncertainties, which lead to a steady-state tracking er-
where esel [k + 1|k] is the predicted state error with the modu- ror. In order to prevent the steady-state tracking error, the state
lation factor μ, and it can be expressed as reference (8) is compensated by integrating the tracking error
esel [k + 1|k] = x∗ (Te∗ , λ∗ ) − Ax [k] − B μ vsel [k] . (32) x∗s [k] = x∗ (Te∗ , λ∗ ) + Ks es [k] (38)

 
BW vsel x∗ T (Te∗ , λ∗ ) − AT xT [k] + vsel
T
B T W (x∗ (Te∗ , λ∗ ) − Ax [k])
μ∗ = T
(33)
2vsel B T W Bvsel

 
∗ BW vsel x∗ T (Te∗ , λ∗ ) − AT xT [k + 1] + vsel
T
B T W (x∗ (Te∗ , λ∗ ) − Ax [k + 1])
μ = T B T W Bv
(37)
2vsel sel
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1921

TABLE II 2) Step 2: Compute the reference rotor flux λ∗ from (14) and
INDUCTION MOTOR AND CONTROL PARAMETERS
(15) using η of step 1. Then, compute the reference state
x∗ (Te∗ , λ∗ ) from (12), block of “computation of reference
Quantity Symbol Value
states.”
DC-bus volt [V] Vd c 450 3) Step 3: Estimate the rotor flux λdqr (k) using the flux
Number of Poles P 4 estimator based on the measured stator current idqs (k)
Rated voltage [V] Y /Δ 220/380
Rated current [A] 14.2/8.2
and rotor speed ωr (k). Compose the current state x[k]
Stator Resistance [Ω] Rs 1.77 using the estimated rotor flux and measured stator current.
Rotor Resistance [Ω] Rr 1.275 Then, predict x[k + 1|k] based on x[k] and the control
Stator Inductance [H] Ls 0.157
Rotor Inductance [H] Lr 0.158
input of previous time step, block of “FCS-MPC with
Mutual Inductance [H] Lm 0.15 time delay compensation.”
Inertia coefficient [Kg.m2 ] J 0.00006 4) Step 4: For each voltage vector listed in Table I, pre-
Rated Motor Speed [rpm] ωN 1740
Rated Power [kW] PN 3.7
dict x[k + 2|k] and evaluate Ji (k) according to (35) and
(39), respectively. Choose the output voltage vector vsel
that yields the minimum cost Ji (k) among the finite volt-
age vectors. Furthermore, compensate the reference states
x∗ (Te∗ , λ∗ ) for the uncertainties by using an integrator as
(38) to yield x∗s [k], block of “cost function minimize.”
5) Step 5: Compute the modulation factor μ∗ from (40) using
vsel of step 4. Finally, apply vsel to the inverter during the
period of μ∗ h. (Note that it is natural to constrain μ∗ as
0 ≤ μ∗ ≤ 1), the block of “duration factor.”

IV. SIMULATION RESULTS


This section provides simulation results for the proposed
FCS-MPC. Table II lists the parameters of the IM used in the
Fig. 2. Break-point torque/speed with flux-increased and flux-limited simulation studies.
operation.

A. Feasibility Analysis of Voltage Vectors and Operating


Regions
where es [k] = es [k − 1] + (x∗ (Te∗ , λ∗ ) − x[k]) is the state er- At each speed, the boundary torque between flux-increased
ror considering the integrator with a proper gain Ks . This com- and limited-flux regions is defined as a break-point torque. A
pensated reference state x∗s [k] substitutes for reference state of relation between the speed and break-point torque can be ob-
the cost index (36) to yield tained offline using the OMS algorithm of Section III-B to build
a look-up table containing break-point torques versus the ro-
Js (k + 2) = (x∗s [k] − x [k + 2|k])T W (x∗s [k] − x [k + 2|k]) .
tational speed. The linear interpolation technique is used dur-
(39)
ing the online running. This look-up table was validated by
Also, this compensated reference state x∗s [k] substitutes for
the simulation data. In constant torque region of the tested
reference state of the modulation factor (37) to yield (40), shown
IM, the two operational regions are described in Fig. 2. This
at the bottom of the page.
shows the normal MTPA in flux-increased and flux-limited re-
Note that it is natural to constrain μ∗ as 0 ≤ μ∗ ≤ 1. The
gions for various break-point torque values at low/high-speed
FCS-MPC procedure is now carried out with J(k) and μ∗ of
operation. The speed and torque boundaries are the rated values
(39) and (40), respectively, subject to (35) and (38).
of the IM under control.
The online procedure of the controller at each time step can
be summarized with the aid of Fig. 1 as follows.
Online procedure of the FCS-MPC algorithm: B. Feasibility Analysis of Control Inputs
1) Step 1: For the given reference torque Te∗ and rotational First of all, feasibility analysis is carried out at the rated torque
speed N ∗ , the operational mode is determined using the of 20 N·m. Fig. 3 compares the steady-state control inputs of
look-up table mentioned in Section III-B. Then η is de- (9) for different settings of η. These are obtained from (20)
fined, i.e., in MTPA mode η is set to be 1 as (20) and with system matrices at different rotating speeds when η = 1
in flux-limited mode η is determined as (21), block of [see Fig. 3(a)], and η was determined as (21) [see Fig. 3(b)].
“MTPA/Flux-limited.” The dotted circle represents the circular boundary of the control

 
∗ BW vsel x∗s T [k] − AT xT [k + 1] + vsel
T
B T W (x∗s [k] − Ax [k + 1])
μ = T B T W Bv
(40)
2vsel sel
1922 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

the same cost index as described in Section III. First, the opti-
mal voltage vector is applied without duration control following
the conventional FCS-MPC as in [2], [5], [6], and [8], which
will be called as conventional MPC. Second, following [12],
the optimal durations are obtained as (40) for all active volt-
age vectors and the combination of voltage vector and duration
that minimizes the cost index is applied to the inverter, which
will be called as MPC-I. The third is the proposed method de-
scribed in the “Online procedure of the FCS-MPC” at the end of
Section III, which will be called as proposed MPC-II. Note that
the proposed method is different from the earlier works in many
aspects, i.e., considering MTPA and flux-limited operation, us-
ing different optimization procedure, providing a systematic
way to determine the weighting matrices. In the comparison,
however, we focus only on the optimization procedure.
Simulations are carried out at different rotational speeds un-
der the assumption that the speed is maintained by external
machines. The torque reference is changed from 0 to 5 to 10 to
15 and to 20 N·m.
The performances at the whole speed range with change of
reference torque are investigated in the simulation. During this
dynamic process, torque, phase current, rotor flux, η, and d-q
currents are observed. The performances at 300 and 1740 r/min
are depicted in Figs. 4 and 5, respectively. The results of the
three methods are depicted in Fig. 4 at low speed of 300 r/min.
Referring to Fig. 4(c), it can be seen that the torque tracks its
reference with lower ripples compared with conventional MPC
and MPC-I shown in Fig. 4(a) and (b). Also, the d-q current
components are of equal value in steady states, which assures
the MTPA operation with η = 1. Rotor flux is increased as the
torque increases due to the increase of ids . Comparing with
conventional MPC and MPC-I, the proposed MPC-II method
minimizes the ripples in phase current, d-q currents and torque
at all torque levels.
When the torque reference is higher than the break-point
torque, the IM is controlled in flux-limited mode to prevent
the saturation of magnetic circuit. The transition from flux-
increased to flux-limited mode can be observed in Fig. 5 at
1740 r/min. Torque/ampere ratio η is decreased from 1 to the
Fig. 3. Steady-state control inputs at different rotational speed with limited value according to (21) at torques of 10 N·m. From the
torque reference 20 N·m. (a) η = 1.0 (20) for flux increased and (b) η for
flux limited (21).
estimated flux trajectory, it can be seen that the value of rotor
flux is varying according to the variations of the torque reference
in flux-increased operation, and then the flux is limited when
input. According to Fig. 3(a), this motor cannot deliver 20-N·m the torque exceeds the break-point value. From the pictures of
output torque at the speed of 1200 r/min with feasible control Figs. 4 and 5, it can be observed that the proposed MPC-II can
inputs when the flux reference is determined with η = 1. But, minimize the torque and current ripples at different torque levels
as shown in Fig. 3(b), 20-N·m output torque can be delivered even at medium and high speeds.
at 1200 r/min and also at 1800 r/min if the flux reference is
determined with η defined by the flux-limited technique. Thus,
we can conclude that the constant torque operating region of V. EXPERIMENTAL RESULTS
the IM can be optimally exploited by adapting the torque per
To verify the performance of the proposed control system, an
ampere ratio η. experimental setup has been established as shown in Fig. 6.
The setup used for experimental validation of the proposed
C. Dynamic Performances With Change of Torque method consists of 3.7 kW IM driven by a 5.6-kW two-level
Reference
voltage-source inverter (VSI) with its interface circuits. The
To validate the effectiveness of the proposed method, com- switching devices in the inverter are insulated-gate bipolar tran-
parisons are made with other two FCS-MPC strategies using sistors (IGBTs) with 20-kHz switching frequency, and the dc-
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1923

Fig. 4. Simulation results: torque, phase current, rotor flux, η, and d-q Fig. 5. Simulation results: torque, phase current, rotor flux, η, and d-q
currents at speed of 300 r/min: (a) conventional MPC, (b) MPC-I, (b) currents at speed of 1740 r/min: (a) conventional MPC, (b) MPC-I, (b)
proposed MPC-II. proposed MPC-II.
1924 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

Fig. 6. Setup of motor–generator (MG) set and two-level inverter for


the experiment.

link voltage is 450 V. The parameters of the motor are presented


in Table II. The experiments are carried out at constant speeds
maintained by 5.5-kW IM driven by a 5.5-kW commercial in-
verter with vector control. The experimental setup incorporates
a control board, which is based on the 32-bit floating point
TMS320F28335 DSP. The current was measured directly using
a current probe, but the torque and flux responses computed by
the embedded controller were transferred to D/A converters and
measured by an oscilloscope. The rotor angle is measured us-
ing standard incremental encoder with 1024 lines. The control
method is implemented with sampling frequency of 10 kHz.
The primary side of the transformer is adjusted at 280 Vac to fix
the dc-link voltage at 450 Vdc.

A. Dynamic State Performance With Change in Torque


Reference
The performances of three MPC methods are investigated at
different speed levels in the experiment. During this dynamic
process, torque, phase current, rotor flux, torque/ampere ratio
η, d-q currents, and d-q voltage components are observed from
top to bottom. The voltage components are recorded during the
experiments, which are related to the optimal cost index and
optimal duration factor at each sampling interval.
Figs. 7 and 8 show the results for a conventional MPC,
reported MPC-I [12], and proposed MPC-II. These figures
show the experimental results at different torque levels of
0, 5, 10, 15, and 20 N·m. At low speed of 300 r/min shown
in Fig. 7, torque tracking is accomplished with high ripples
using conventional MPC [see Fig. 7(a)], low ripples with the
MPC-I [see Fig. 7 (b)] and very low ripples with the proposed
MPC-II [see Fig. 7 (c)]. It can be seen that d-q currents are
equal at steady state that assures the MTPA operation, in which
torque/ampere ratio η is set to be 1. Rotor flux is also seen to be
proportional to torque values. Comparing to the conventional
MPC and MPC-I, the proposed MPC-II minimizes the ripples
in phase current, d-q currents, and torque at all torque levels.
When the operating point (i.e., torque/speed) exceeds the
Fig. 7. Experimental results: torque, phase current, rotor flux, η, and
break-point boundary, the IM is controlled in flux-limited mode. d-q current at 300 r/min: (a) conventional MPC, (b) MPC-I, (c) proposed
The transition from flux-increased to flux-limited mode is shown MPC-II.
in Fig. 8 at rated speed of 1740 r/min. As explained in Section III
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1925

and depicted in simulations, η is decreased from 1 to the limited


value (21) at torque of 10 in Fig. 8. Also, it can be noticed that
the rotor flux is varying according to variations of the torque ref-
erence value in flux-increased operation, and then the flux is lim-
ited when the torque/speed point exceeds the break-point value.
Also, from the experimental pictures in Figs. 7 and 8, it can
be observed that the proposed MPC-II can minimize the torque
and current ripples at different torque levels even at medium and
high speeds. Also, voltage components are increased with the
increasing speed. From close observation, it can be noticed that
the proposed MPC-II will deliver torque and current with less
ripples than that of conventional MPC and MPC-I.

B. Steady-State Performance
Figs. 9 and 10 show the steady-state performance of the IM
at 1200 r/min and rated torque of 20 N·m for the conventional
MPC, MPC-I, and proposed MPC-II in simulation and experi-
ments, respectively. From a close observation of the simulation
results shown in Fig. 9, it can be noticed that the torque error is 3,
1.5, and 1.0 N·m for conventional MPC, MPC-I, and proposed
MPC-II, respectively. From a close observation of experimen-
tal results shown in Fig. 10, it can be noticed that the torque
error is 5, 2, and 1.3 N·m for conventional MPC, MPC-I, and
proposed MPC-II, respectively, which are matched to the sim-
ulation results. Some glitches are emerged to the experimental
results, especially in torque signal, which are mainly caused by
the noise and limited resolution of A/D or/and D/A converters.
Nevertheless, it is still clear to observe the improved perfor-
mance of the proposed method. Also, it is seen that both MPC-I
and proposed MPC-II present much lower current oscillations
than that of conventional MPC, especially in MPC-II. It can be
concluded here that the torque ripples are reduced considerably
by MPC-II.
Fig. 11 shows phase current and its spectrum at 300 r/min
and rated torque of 20 N·m. The current wave is improved us-
ing the proposed MPC-II as shown in Fig. 11(c). Also, current
spectrum shown in Fig. 11 shows that the fundamental wave is
pointed at 12 Hz (300r/min) and some harmonics are pointed
at high frequencies. From close observation, it can be noticed
that the harmonics are reduced with the proposed MPC-II com-
pared with other methods. In comparison with the simulation
results, it is worthwhile to mention here that the flux-limited al-
gorithm enables the IM to work in real world without saturation.
Moreover, the magnetizing inductance Lm could be considered
constant for practical operation (real motor without saturation)
as well considered in the simulation model. Hence, the presented
MTPA in both operating regions is valid.

C. Acceleration Test
Torque tracking in real IM is proven in an acceleration test, as
shown in Fig. 12 for the three MPC methods. In Fig. 12(a), the
performance of the conventional MPC is shown for step changes
in the reference torque. In t = 0.5 s, a torque reference of 6 N·m
Fig. 8. Experimental results: torque, phase current, rotor flux, η, and is used and the machine accelerates to 1700 r/min, later the
d-q current at 1740 r/min: (a) conventional MPC, (b) MPC-I, (c) proposed
MPC-II. reference is set again to zero. The same maneuver was carried
out for MPC-I and proposed MPC-II [see Fig. 12(b) and (c)].
1926 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

Fig. 9. (Simulation results): Torque error at steady state at 1200 r/min and torque of 20 N·m: (a) Conventional MPC, (b) MPC-I, (c) proposed
MPC-II.

Fig. 10. (Experimental results): Torque error at steady state at 1200 r/min and torque of 20 N·m: (a) conventional MPC, (b) MPC-I, (c) proposed
MPC-II.

Fig. 11. (Experimental results): Current and frequency spectrum at steady state at 300 r/min and torque of 20 N·m: (a) conventional MPC, (b)
MPC-I, (c) proposed MPC-II.

Fig. 12. Acceleration test (experimental results): (a) conventional MPC, (b) MPC-I, (c) proposed MPC-II.

As shown in these torque waveforms, a good tracking (with low increased property was applied in high torque at low speeds and
overshoot) of the torque reference is achieved by the proposed in low torque at high speed. The flux-limited technique was suc-
MPC-II. cessfully applied in high speeds at high torque levels. Improve-
ments in the performance were observed when the new control
system was compared with the reported results. The contribu-
VI. CONCLUSION tion here is that the proposed MPC-II exhibits the lowest ripple
In the proposed MPC, a method to determine the reference torque and smallest harmonics in stator current at all torque
states was presented so that the torque reference tracking is ob- and speed levels. In addition, as it is shown from simulation
tained while enabling the MTPA control in flux increased region and experimental output, fast dynamic response was achieved
and flux limited region for whole torque levels. MTPA with flux- with hard-step torque reference change. Also, different weight-
AHMED et al.: FCS-MPC METHOD FOR TORQUE CONTROL OF IMS USING A STATE TRACKING COST INDEX 1927

ing factors were used depending on the rotational speeds. The [21] S. Dymko, S. Peresada, and R. Leidhold, “Torque control of saturated
proposed scheme was found to be very promising and valuable induction motors with torque per ampere ratio maximization,” in Proc.
IEEE Int. Conf. Intell. Energy Power Syst., Kiev, Ukraine, Jun., 2014, pp.
as compared with the reported MPC-I and conventional MPC. 251–256.
[22] J. Holtz, “The dynamic representation of AC drive systems by complex
REFERENCES signal flow graphs,” in Proc. IEEE Int. Symp. Ind. Electron., Santiago,
Chile, May 1994, pp. 1–6.
[1] M. Nemec, D. Nedeljkovic, and V. Ambrozic, “Predictive torque control [23] N. P. Quang and J. A. Dittrich, “Machine models as prerequisite to de-
of induction machines using immediate flux control,” IEEE Trans. Ind. sign the controllers and observers,” in Vector Control of Three-Phase AC
Electron., vol. 54, no. 4, pp. 2009–2017, Aug. 2007. Machines. Berlin, Germany: Springer-Verlag, 2008, pp. 61–105.
[2] H. Miranda, P. Cortes, J. Yuz, and J. Rodriguez, “Predictive torque control [24] P. Cortes, M. P. Kazmierkowski, R. M. Kennel, D. E. Quevedo, and J.
of induction machines based on state-space models,” IEEE Trans. Ind. Rodriguez, “Predictive control in power electronics and drives,” IEEE
Electron., vol. 56, no. 6, pp. 1916–1924, Jun. 2009. Trans. Ind. Electron., vol. 55, no. 12, pp. 4312–4324, Dec. 2008.
[3] T. Geyer, G. Papafotiou, and M. Morari, “Model predictive direct torque [25] G. Dong and O. Ojo, “Efficiency optimizing control of induction mo-
control—Part I: Concept, algorithm, and analysis,” IEEE Trans. Ind. Elec- tor using natural variables,” IEEE Trans. Ind. Electron., vol. 53, no. 6,
tron., vol. 56, no. 6, pp. 1894–1905, Jun. 2009. pp. 1791–1798, Dec. 2006.
[4] S. K. Kim, J. S. Kim, and Y. L. Lee, “Model predictive control based direct [26] M. Cacciato, A. Consoli, G. Scarcella, G. Scelba, and A. Testa, “Efficiency
torque control of permanent magnet synchronous motors,” in Proc IEEE optimization techniques via constant optimal slip control of induction
Int. Symp. Ind. Electron., Taipei, Taiwan, May 2013, pp. 1–6. motor drives,” in Proc. Int. Symp. Power Electron., Elect. Drives, Autom.
[5] F. Wang, S. Li, X. Mei, W. Xie, J. Rodrı́guez, and R. M. Kennel, “Model- Motion Sicily, Italy, May 2006, pp. 33–38.
based predictive direct control strategies for electrical drives: An experi- [27] S. K. Kim, H. S. Park, and Y. I. Lee, “Stabilizing model predictive control
mental evaluation of PTC and PCC methods,” IEEE Trans. Ind. Informat., for torque control of permanent magnet synchronous motor,” in Proc. 33rd
vol. 11, no. 3, pp. 671–681, Jun. 2015. Chin. Control Conf., Dec. 2013, pp. 7772–7777.
[6] C. S. Lim, E. Levi, M. Jones, N. A. Rahim, and W. P. Hew, “FCS-MPC- [28] M. S. Cavalca, R. K. H. Galvao, and T. Yoneyama, “Robust model pre-
based current control of a five-phase induction motor and its compari- dictive control using linear matrix inequalities for the treatment of asym-
son with PI-PWM control,” IEEE Trans. Ind. Electron., vol. 61, no. 1, metric output constraints,” J. Control Sci. Eng., vol. 2012, Jan. 2012,
pp. 149–163, Jan. 2014. Art. no. 485784.
[7] E. Fuentes, J. Rodriguez, C. Silva, S. Diaz, and D. Quevedo, “Speed [29] S. Carpiuc and C. Lazar, “Lyapunov–based constrained explicit current
control of a permanent magnet synchronous motor using predictive current predictive control in permanent magnet synchronous machine drives,” in
control,” in Proc. 6th IEEE Int. Power Electron. Motion Control Conf. Proc. Int. Symp. Power Electron., Elect. Drives, Autom. Motion Ischia,
Wuhan, China, May 2009, pp. 390–395. Italy, Jun., 2014, pp. 461–466.
[8] A. A. Ahmed, “Fast-speed drives for permanent magnet synchronous [30] M. Preindl, “Robust control invariant sets and lyapunov-based MPC for
motor based on model predictive control,” in Proc. IEEE Veh. Power IPM synchronous motor drives,” IEEE Trans. Ind. Electron., vol. 56, no. 6,
Propulsion Conf. Montréal, QC, Canada, Oct. 2015, pp. 1–6. pp. 1894–1905, Jun. 2009.
[9] I. Takahashi and T. Noguchi, “A new quick-response and high-efficiency [31] J. Rodriguez and P. Cortes, Predictive Control of Power Converters and
control strategy of an induction motor,” IEEE Trans. Ind. Appl., vol. 22, Electrical Drives. Hoboken, NJ, USA: Wiley, 2012.
no. 5, pp. 820–827, Sep./Oct. 1986.
[10] Y. Zhang and H. Yang, “Torque ripple reduction of model predictive torque
control of induction motor drives,” in Proc. Energy Convers. Congr. Expo.
Denver, CO, USA, Sep. 2013, pp. 1176–1183.
[11] P. Karamanakos, P. Stolze, R. M. Kennel, S. Manias, and H. T. Mou- Abdelsalam A. Ahmed (M’16) was born in
ton, “Variable switching point predictive torque control of induction Kafrelsheikh, Egypt. He received the M.S. de-
machines,” IEEE J. Emerg. Sel. Topics Power Electron., vol. 2, no. 2, gree in electrical engineering from Tanta Univer-
pp. 285–295, Jun. 2014. sity, Tanta, Egypt, in 2008, and the Ph.D. degree
[12] Y. Zhang and H. Yang, “Model predictive torque control of induction motor in electrical engineering and automation from
drives with optimal duty cycle control,” IEEE Trans. Power Electron., Harbin Institute of Technology, Harbin, China, in
vol. 29, no. 12, pp. 6593–6603, 2014. 2012.
[13] Y. Zhang and H. Yang, “Model-predictive flux control of induction motor Upon completion of the Ph.D. degree, he
drives with switching instant optimization,” IEEE Trans. Energy Convers., joined the Department of Electrical Power and
vol. 30, no. 3, pp. 1113–1122, Sep. 2015. Machines Engineering, Faculty of Engineering,
[14] Y. Cho, K. B. Lee, J. H. Song, and Y. I. Lee, “Torque-ripple minimization Tanta University, where he is currently an Assis-
and fast dynamic scheme for torque predictive control of permanent- tant Professor. From July 2013 to July 2014, he was a Postdoctoral Fel-
magnet synchronous motors,” IEEE Trans. Power Electron., vol. 30, no. 4, lowsh in the Department of Instrumental Science and Technology, School
pp. 2182–2190, Apr. 2015. of Electrical Engineering and Automation, Harbin Institute of Technology.
[15] M. Habibullah, D. D. C. Lu, D. Xiao, and M. F. Rahman, “A simplified From September 2015 to September 2016, he was a Postdoctoral Fellow
finite-state predictive direct torque control for induction motor drive,” in the Department of Electrical and Information Engineering, Seoul Na-
IEEE Trans. Ind. Electron., vol. 63, no. 6, pp. 3964–3975, Jun. 2016. tional University of Science and Technology, Seoul, South Korea. He has
[16] M. Preindl and S. Bolognani, “Model predictive direct torque control authored more than 20 scientific papers in the field of modern strategies
with finite control set for PMSM drive systems—Part 1: Maximum of control and drives of electric machines, electric and hybrid electric
torque per ampere operation,” IEEE Trans. Ind. Informat., vol. 9, no. 4, vehicles, and power electronics. His current research interests include
pp. 1912–1921, Nov. 2013. model predictive control in electrical drive systems and power converters
[17] O. Wasynczuk et al., “A maximum torque per ampere control strategy and electric vehicle drive systems.
for induction motor drives,” IEEE Trans. Energy Convers., vol. 13, no. 2,
pp. 163–169, Jun. 1998.
[18] R. Bojoi, Z. Li, S.A. Odhano, G. Griva, and A. Tenconi, “Unified direct-
flux vector control of induction motor drives with maximum torque per Byung Kwon Koh was born in South Korea
ampere operation,” in Proc. Energy Convers. Congr. Expo. Denver, CO, in 1990. He received the B.S. degree in elec-
USA, Sep. 2013, pp. 3888–3895. trical and information engineering from Seoul
[19] C. Kwon and S. D. Sudhoff, “An improved maximum torque per amp National University of Science and Technology,
control strategy for induction machine drives,” in Proc. 12th Annu. IEEE Seoul, South Korea, in 2015, where he is cur-
Appl. Power Electron. Conf. Expo. Austin, TX, USA, Mar., 2005, pp. 740– rently working toward the M.S. degree in electri-
745. cal and information engineering.
[20] J. H. Mun, J. S. Ko, J. S. Choi, M. G. Jang, and D. H. Chung, “Maximum His research interests include electric ma-
torque control of IM drive using AIPI controller,” in Proc. Int. Conf. chine drives and control engineering.
Control Autom.. Syst., Gyunggi-do, South Korea, Oct., 2010, pp. 1223–
1228.
1928 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 64, NO. 3, MARCH 2017

Hyo Sung Park was born in South Korea in Young Il Lee (M’99–SM’15) received the B.Sc.,
1987. He received the B.S. and M.S. degrees M.S., and Ph.D. degrees in control and instru-
in control and instrumentation engineering from mentation from Seoul National University, Seoul,
Seoul National University of Science and Tech- South Korea, in 1986, 1988, and 1993, respec-
nology, Seoul, South Korea, in 2013 and 2015, tively.
respectively. From 1994 to 2001, he was an Associate
He is currently with LG Electronics, Seoul, Professor at Gyeongsang National University.
South Korea. His research interests include He has been with Seoul National University of
electric machine drives and control engineering. Science and Technology, Seoul, since 2001 as
a Professor. From 1998 to 1999 and in 2007,
he was with Oxford University as a Visiting Re-
search Fellow. His research interests include model predictive control
and its application to power converters, electrical machines, and electric
vehicles.
Dr. Lee was the Editor of the International Journal of Control, Automa-
tion and Systems.
Kyo-Beum Lee (S’02–M’04–SM’10) received
the B.S. and M.S. degrees in electrical and elec-
tronic engineering from Ajou University, Suwon,
South Korea, in 1997 and 1999, respectively,
and the Ph.D. degree in electrical engineering
from Korea University, Seoul, South Korea in
2003.
From 2003 to 2006, he was with the Insti-
tute of Energy Technology, Aalborg University,
Aalborg, Denmark. From 2006 to 2007, he was
with the Department of Electrical Engineering,
Chonbuk National University, Jeonju, South Korea. In 2007, he joined
the Department of Electrical and Computer Engineering, Ajou Univer-
sity. His research interests include electric machine drives, renewable
power generation, and electric vehicle applications.
Dr. Lee is an Associate Editor of the IEEE TRANSACTIONS ON POWER
ELECTRONICS, The Journal of Power Electronics, and The Journal of
Electrical Engineering and Technology.

You might also like