You are on page 1of 10

4888

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 58, NO. 10, OCTOBER 2011

Sliding Mode Control of a Stand-Alone Wound


Rotor Synchronous Generator
Ral Santiago Muoz-Aguilar, Arnau Dria-Cerezo, Enric Fossas, Member, IEEE, and Rafel Cardoner Papal

AbstractThis paper presents a sliding mode control for a


wound rotor synchronous machine acting as an isolated generator.
The standard dq model of the machine is connected to a resistive
load. A switching function is defined in order to fulfill control
objectives, and the ideal sliding dynamics is proved to be stable.
From the desired surface, the standard sliding methodology is
applied to obtain a robust and very simple controller. Numerical
simulations and experimental results validate the control law and
show good performance and a fast response to load and reference
changes.
Index TermsElectrical machines, sliding mode control (SMC),
stand-alone generation, stator voltage amplitude regulation,
wound rotor synchronous generator (WRSM).

I. I NTRODUCTION

LECTRICAL energy has been extensively studied due


to its transmission possibilities, flexibility, and control
capacity, among others. The energy demand, as well as interest
in a clean efficient generation, is increasing worldwide. Electrical energy is mainly generated by interconnecting wound
rotor synchronous machines (SMs) (WRSMs) which set up a
theoretically infinite bus. Hence, WRSMs are generally studied
connected to an infinite bus called power grid [1].
When the WRSM is connected to the power grid, the latter
determines the stator voltage and frequency while the WRSM,
by means of the rotor voltage, helps to improve the power factor
and compensate the reactive power at the connection point.
Usually, the stator transients can be neglected [2].
A significantly different scenario is when the WRSM is
isolated from the grid. This paper focuses on the stand-alone
case of the WRSM where neither the stator amplitude nor the

Manuscript received December 1, 2009; revised March 31, 2010 and


October 6, 2010; accepted January 25, 2011. Date of publication February 17,
2011; date of current version August 30, 2011. This work was supported in part
by the Spanish Government research projects ENE2008-06841-C02-01/ALT
(R. S. Muoz-Aguilar), DPI2010-15110 (A. Dria-Cerezo and E. Fossas), and
DPI2008-01408 (E. Fossas).
R. S. Muoz-Aguilar is with the Department of Electrical Engineering, Universitat Politcnica de Catalunya, 08222 Terrassa, Spain (e-mail: raul.munozaguilar@upc.edu).
A. Dria-Cerezo is with the Department of Electrical Engineering and the
Institute of Industrial and Control Engineering, Universitat Politcnica de
Catalunya, 08800 Vilanova i la Geltr, Spain (e-mail: arnau.doria@upc.edu).
E. Fossas is with the Department of Automatic Control and the Institute
of Industrial and Control Engineering, Universitat Politcnica de Catalunya,
08028 Barcelona, Spain (e-mail: enric.fossas@upc.edu).
R. Cardoner Papal is with the Institute of Industrial and Control Engineering, Universitat Politcnica de Catalunya, 08028 Barcelona, Spain (e-mail:
rafel.cardoner@upc.edu).
Color versions of one or more of the figures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TIE.2011.2116754

frequency is fixed. In this configuration, the mechanical speed


determines the frequency and voltage amplitude at the same
time, and the rotor voltage sets the stator voltage amplitude.
The SM can be controlled by several techniques. In industry,
the most common are linear techniques [3], [4] and hysteresis
regulators [5], [6]. However, decoupling methods [7], which are
widely employed for asynchronous machines, are also extended
to the synchronous case. Modern control techniques, such as
passivity-based control [8], optimal torque control [9], [10], or
predictive control [11], are also used for SM speed regulation.
In this paper, the WRSM is regulated in the sliding mode
control framework. Its robustness and ease of implementation
make this approach particularly attractive [12]. In general, the
sliding mode control is suitable for variable structure systems
(VSS), such as power converters [13]. Even though electrical
machines are not VSS, sliding mode control has been suggested
as appropriate for their control [14] mainly because of the use of
power converters when applying the electrical machine control
voltages and the discrete values taken by the control voltage.
The combination of sliding mode control and singular perturbation theory [15] and the former and the block control
approach [16] has been proposed for a WRSM connected to the
power grid. Other examples of sliding techniques can be found
in multimachine systems [17] and motor applications [18], [19].
This methodology was already used for a generator in a standalone configuration in [20].
The main contribution of this paper is a sliding mode control
algorithm for a stand-alone wound rotor synchronous generator with a resistive load. Contrary to [20], where the whole
state and the machine parameters are required, the obtained
controller is robust to variations in the machine and load
parameters, only requires voltage measurements, and is easily
implementable. The control law is experimentally tested, and
it performs well and responses rapidly. Moreover, the local
stability of the closed loop dynamics is proved through a smallsignal model.
II. S YSTEM D ESCRIPTION
Fig. 1 shows the following scenario: A primary mover (for
example, an internal combustion engine, a hydroturbine, etc.)
drags a WRSM at a constant speed. The WRSM, in turn,
acting as a generator feeds an isolated load. In this figure
and throughout this paper, the convention of incoming positive
current is adopted for both stator and field currents.
As explained before, this system differs from the typical
grid connection. In an isolated configuration, the frequency is
determined by the mechanical speed m which is provided by

0278-0046/$26.00 2011 IEEE

MUOZ-AGUILAR et al.: SLIDING MODE CONTROL OF A STAND-ALONE WOUND ROTOR SYNCHRONOUS GENERATOR

Fig. 1.

4889

Diagram of a stand-alone wound rotor synchronous generator.


Fig. 2. Equilibrium point distribution.

the primary source and externally regulated while the voltage


amplitude is set by the rotor field voltage.
A. Dynamic Model
From the well-known dynamic equations (in dq-coordinates)
of the WRSM and the interconnection rules with a pure resistive
load, the whole dynamic system is presented.
The electrical part of the WRSM1 is given by


Rs Ls
vd
0
dx
= Ls Rs Lm x + vq (1)
L
dt
0
0
RF
vF

and


0
B = 0.
1

B. Control Objective

where

Ls
L= 0
Lm

0
Ls
0

Lm
0
LF

is the inductance matrix; xT = (id , iq , iF ) R3 are the dq


stator and field currents, respectively; Rs and RF are the
stator and field resistances; Ls , Lm , and LF are the stator,
magnetizing, and field inductances; is the electrical speed
( = np m , where np is the number of pole pairs); vd and vq
are the dq stator voltages; and vF is the field voltage which will
be used as a control input.
In order to design the control law, let us first obtain the
complete model of a WRSM connected to a resistive load
RL . The interconnection diagram is depicted in Fig. 1, where
T
= (vLd , vLq ) R2 and iTL = (iLd , iLq ) R2 are the load
vL
voltages and currents in dq-coordinates related by
 
 
iLd
vLd
= RL
.
(2)
vLq
iLq

vs = vL
iL = is
where vsT = (vd , vq ) R2 and iTs = (id , iq ) R2 are the stator
voltages and currents in dq-coordinates. Now, putting together
(1) and (2), the system can be written in an affine form as
L

dx
= Ax + BvF
dt

cylindrical rotor type without damping windings is considered.

As already mentioned, for a stand-alone SM, the stator


frequency is directly given by the mechanical speed, which, in
this paper, is assumed to be constant and externally regulated.
Thus, the system output is the stator voltage amplitude Vs ,
which must be regulated by the rotor voltage vF . The stator
voltage can be easily obtained in dq-coordinates as

Vs = vd2 + vq2
(4)
or using (2) as a function of currents

Vs = RL i2d + i2q .

(5)

C. Equilibrium Points
From (3) and (5), where Vs was replaced by its desired value
Vref , the equilibrium point (id , iq , iF ) and the control input vF
must fulfill
0 = (Rs + RL )id + Ls iq

According to Fig. 1, the interconnection rules are

1A

where L is the inductance matrix defined earlier

(Rs + RL )Ls
0
A=
(Rs + RL ) Lm
Ls
00 RF

(3)

0 = Ls id (Rs
0 = RF iF + vF
2

2
2
id + i2
.
= RL
Vref
q

RL )iq

(6)

Lm iF

(7)
(8)
(9)

Note that (6) and (9) can be interpreted as the intersection of


a cylinder and a straight line (see Fig. 2).
Using cylindrical coordinates (Is , , iF ) where
id = Is cos

(10)

iq = Is sin

(11)

4890

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 58, NO. 10, OCTOBER 2011

(6) and (9) result in


Is =

By replacing the matrices and partial derivative values in


(17), we have

Vs
RL

0=

ueq = RF iF

Rs + RL
Ls
Vref cos +
Vref sin .
RL
RL

Then, the value2 in equilibrium is




Rs + RL
= arctan
.
Ls

(12)

From (7) and (9)(12), the field current equilibrium value is


iF =

Vref
(Ls cos + (Rs + RL ) sin ) .
Lm RL

(13)

Finally, using (8), the equilibrium control input is obtained


vF =

RF
Vref (Ls cos + (Rs + RL ) sin ) .
Lm RL

(14)

LF
(Rs + RL )id + Lm iq
Lm


iq
(Rs + RL )

iq + iF
Ls
Lm
id

(18)

where = Ls LF L2m , which is always positive.


In the 3-D space (id , iq , iF ), the sliding surface is a cylinder.
Sliding motion can be expected only in the cylinder subset
defined by id = 0 where the transversality condition holds. The
closer id is to zero, the higher the equivalent control. Therefore,
a closed subset of the sliding domain must be taken to obtain a
bounded control effort ueq .
As expected, from (14), vF = ueq (id , iq , iF ).
B. Sliding Mode Controller
The sliding mode controller must ensure that

III. C ONTROL D ESIGN

In this section, the sliding mode control technique is applied


to regulate a stand-alone wound rotor synchronous generator.
The switching function is directly derived from the stator voltage amplitude, and assuming a bangbang control action, the
switching policy is defined to guarantee sliding modes. Finally,
the equivalent control is used to prove the local stability of the
ideal sliding dynamics (ISD).
A. Switching Function and Equivalent Control
According to the control goals, we define the switching
function s(x) as follows:
2
s(x) = Vs2 Vref
.

(15)

Note that the switching function contains the square error


of the voltage amplitude Vs2 instead of the classical error
difference. As shown hereinafter, this choice implies that both
equilibrium points are achievable. Using (5), the switching
function can be written in state variables
2

2
2
id + i2q Vref
s(x) = RL
.
(16)
From (3), the equivalent control (i.e., the fictitious control
making the sliding surface flow invariant) fulfills
s 1
L (Ax + Bueq ) = 0.
x
Hence

ueq =

s 1
L B
x

1

s 1
L Ax.
x

(17)

2 As the field magnetomotive force is on the q-axis, this angle can be


seen as the load angle, which is widely studied for a grid connection to ensure
synchronism of the WRSM. However, in this case of isolated generation,
synchronism is not an issue.

ds
<0
dt

which is equivalent to
s

s 1
L (Ax + BvF ) < 0.
x

Adding and subtracting Bueq and considering (17)


s

s 1
L B(vF ueq ) < 0.
x

Therefore, the control action defined by




s 1
vF = ueq k sign s L B
x

(19)

(20)

with a positive k, fulfills the stability condition s(ds/dt) < 0




s 1
ds

s
= k s L B 0.
dt
x
Evaluating (s(s/x)L1 B) in (20) and taking into account
2
Lm /(Ls LF L2m )) > 0, (20) can be simplified as
that (2RL
vF = ueq k sign(sid ).

(21)

C. Switching Control Policy


In this system, the control action vF is implemented using a
dcdc power converter which commutes between two discrete
signal values, Vdc and Vdc . Thus, the control law proposed in
(21) must be modified as follows.
From (21), the stability condition simplifies to
sid (vF ueq ) < 0.
Hence, the control rotor voltage

Vdc ,
if sid < 0
vF =
Vdc , if sid > 0

(22)

(23)

MUOZ-AGUILAR et al.: SLIDING MODE CONTROL OF A STAND-ALONE WOUND ROTOR SYNCHRONOUS GENERATOR

Fig. 3.

4891

Control scheme.

provides sliding modes in the subset of s = 0 defined by


Vdc < ueq < Vdc .
For robustness, since voltages are accessible variables and
are used to compute the switching function, the switching
policy in (23) is given in voltage terms. Using (2) in (23),
we get

vF =

Vdc ,
if svd < 0
Vdc , if svd > 0.

(24)

The proposed control scheme is depicted in Fig. 3. is the


rotor position (required to compute the dq-transformation), and
vabc are the three-phase stator voltages. A hysteresis block
is added to limit the switching frequency. It is worth noting that this control action depends on the voltage measurements and rotor position only. Therefore, since the switching
function, written as a function of the voltages, is parameter
independent, the closed loop system is robust provided that
Vdc < ueq < Vdc .
It is worth to mention the simplicity of (24) compared with
the control law presented in [20].

D. ISD
The ISD or zero dynamics, i.e., the dynamics defined on
s = 0 by (3) and vF = ueq , will be given in local variables
that parameterize the cylinder, namely, iq and iF . Notice that
this dynamics is well defined on the subset of s = 0 where the
transversality condition holds.
This subset has two connected components, and there is
a symmetry between each component dynamics. From now
on,only the dynamics in the component defined by id =
2 /R2 ) i2 will be considered.
+ (Vref
q
L
Solving (16) for id and putting it in (18) and both in (3), the
ISD is given by (16) and

Fig. 4. State space: Vector field and trajectory for a given initial condition,
iq (0) = 0 and iF (0) = 0.
2
2
Vref
/RL
. Then, (25) and (26) simplify in

diq
a
= Y 2 i2q aiq iF ,
dt
c
1
diF
=
(cY 2 + iq iF ).
dt
2
2
Y i

(27)
(28)

There is a unique equilibrium point (iq , iF ) in this connected


component at



cos
sin
+
(id , iF ) = Y sin , cY
.
a

It can be proved to be locally asymptotically stable by smallsignal analysis (linearization around the equilibrium point).
Indeed, the ISD Jacobian at the equilibrium point is


0
ac
JacISD =
.
(29)
ac cos12 tan
Hence, local stability comes from the determinant
det(JacISD ) =

2
cos2

which is positive and the trace of JacISD

2
Vref
Rs + RL
Lm
2
iq
iF
2 iq
RL
Ls
Ls


2
Rs + RL Vref
1
diF
=
.
2 + iq iF
2
dt
Lm
RL
Vref
2

i
q
R2

diq
=
dt

(25)
(26)

For easy reading, let us define the following positive constants a = (Rs + RL )/Ls , c = (Rs + RL )/Lm , and Y 2 =

tr(JacISD ) = tan =

Rs + RL
Ls

which is negative.
Numerical analysis can also be done. Fig. 4 shows the phase
portrait of the dynamic system defined by (25) and (26) in the
subset (id , iq , iF ) R3 such that

2
Vref
Vref
Vref
2
id = +
< iq <
2 iq and R
RL
RL
L

4892

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 58, NO. 10, OCTOBER 2011

the power grid with a diode rectifier, an L filter, and a capacitor


dc bus. For the experimental tests, the bus voltage is set to
Vdc = 137.5 V.
The resistive load is composed of two interconnected banks,
with half- or full-load values (equivalent to RL = 128 and
RL = 64 for the nominal voltage).

B. Control Implementation
Fig. 5. Hardware interconnection.
TABLE I
WRSM DATA C HARACTERISTICS AND PARAMETERS

and the trajectory through iq (0) = 0, and iF (0) = 0. The


parameter values are those in Section V, where numerical
simulations and experiments are performed. The equilibrium
point, computed from (11) and (13), is iq = 2.41 A and
iF = 10.06 A.3
Notice the local stability of the equilibrium point. However, trajectories starting close to iq = (Vref /RL ) and iF >
(Vref (Rs + RL ))/Lm RL and to iq = Vref /RL and iF <
(Vref (Rs + RL ))/Lm RL escape from the cylinder. An exhaustive analysis of the complete nonlinear dynamics is left as
future work.
IV. H ARDWARE D ESCRIPTION AND
C ONTROLLER I MPLEMENTATION
A. Hardware Description
The experimental setup scheme is shown in Fig. 5. Note that,
in this case, the WRSM is dragged by a dc motor (which emulates the primary mover proposed in Section II). The WRSM
rotor position and stator voltages are measured, the latter using
two differential sensors. These measurements are acquired by
the digital signal processor (DSP), which is programmed from a
personal computer. The dc/dc converter is connected to the grid
with a diode rectifier and a capacitor. This converter applies the
control defined by the DSP card to the WRSM. The dc motor,
which provides the WRSM with constant speed, is a 3-kW
1500-r/min machine with the 4Q2 commercial speed controller
from Control Techniques Drives Ltd.
The WRSM is a 2.4-kVA four-pole three-phase machine.
Nominal characteristics and parameters obtained using IEEE
Std. 115-1995 [21] are in Table I.4
The power converter is a full bridge dc/dc converter which
can provide Vdc voltages. The Vdc voltage is obtained from
3 In this analysis, the rotor parameters are referred to the stator; in the real
application, the rotor current applied to the machine at the equilibrium point
becomes 2.515 A.
4 The apostrophe indicates that the parameters are referred from the rotor to
the stator and that n is the transformation relationship.

The control algorithm is programmed into a Texas Instruments floating-point 150-MHz DSP (DSP TMS 320F28335).
Real-Time Workshop C-code generation from Matlab/
Simulink allows the code implementation to be simplified to
the DSP without using a C-code editor directly. A Texas Target support package is used to configure the analog-to-digital
converter, pulsewidth modulation, serial peripheral interface,
general-purpose inputoutput ports, and interruptions.
The sliding mode controller has been implemented so that a
maximum 10-kHz switching frequency is allowed.

V. S IMULATION AND E XPERIMENTAL R ESULTS


In this section, a set of simulations and experiments using
the controller in Section III is presented. The mechanical speed
is fixed at 1500 r/min (which corresponds to a 50-Hz stator
frequency) by the dc motor.
The simulations were performed by a cosimulation procedure with Matlab/Simulink and Psim. Thus, the same block
controller, coded in Matlab/Simulink, can be used for both
simulations and experimental tests. This procedure provides
simulation results in close agreement with actual system performance, which is really helpful for implementation purposes.
Practically all hardware of the real plant, i.e., dc/dc power
converter, acquisition stages, etc., is built in Psim for simulation
purposes. The WRSM was built using the Matlab Power System
Blockset. According to the sample time programmed in the
DSP, a time step of 104 s is used in Matlab while a time
step of 106 s is used in Psim. For cosimulation purposes, the
integration algorithms are Newton (ode14x) in Matlab and the
trapezoidal rule in Psim.
In the first experiment, the reference line
voltage is set to
380 Vrms (which corresponds to Vs = 220 2 V), and the load
is suddenly increased from the one-half- to full-load value.
Simulation results are summarized in Figs. 68. Fig. 6 shows
the three-phase line stator voltages (vabc ) and the switching
function. The load change occurs at t = 0.5 s. At the bottom
of the figure is a zoom of the top picture around t = 0.5 s. Note
that the sliding mode recovers very fast and the perturbation
effect on abc-voltages lasts less than one period. This rapid
response can also be observed in the stator voltage amplitude
(see Fig. 7).
Simulation results of the actual and equivalent controls are
displayed in Fig. 8. The top picture is zoomed around t = 0.5 s
at the bottom. The actual control takes the discrete values
137.5 V. Note that the average value changes for the new
state at a full-load value. As can be seen, the equivalent control
remains in the expected strip (Vdc < ueq < Vdc ). Fig. 8 also

MUOZ-AGUILAR et al.: SLIDING MODE CONTROL OF A STAND-ALONE WOUND ROTOR SYNCHRONOUS GENERATOR

4893

Fig. 9. Experimental results: Three-phase stator voltages and switching function for a change from the one-half- to full-load value.
Fig. 6. Simulation results: Three-phase stator voltages and switching function
for a change from the one-half- to full-load value.

Fig. 10. Experimental results: Stator voltage amplitude and its reference and
switching control police and its filtered value for a change from the half- to
full-load value.

Fig. 7. Simulation results: Stator voltage amplitude Vs for a change from the
half- to full-load value.

Fig. 8. Simulation results: Field voltage vF and equivalent control ueq for a
change from the one-half- to full-load value.

contains the dcdc switch driver signals, which lie between zero
and one and are equivalent to the applied field voltage.
The experimental results for the three-phase stator voltages
and switching function are shown in Fig. 9. The simulation
and experimental results are in good agreement. In both cases,
the stator voltage amplitude is perfectly regulated, with a fast
time response. The controller needs less than one stator voltage
cycle to recover the reference (details are zoomed at the bottom
of Fig. 9). The switching function oscillates around zero, but
the resulting chattering phenomenon is not reflected in the
experimental test due to the filter effect of the digital-to-analog
converter. Simulation results show that chattering is less than
2%3% of the stator voltage amplitude.
The stator voltage amplitude, reference voltage, and actual
and equivalent controls are depicted in Fig. 10. Note that
experimental results reveal very good regulation performance.
In the second experiment, the stator voltage amplitude reference is changed. From an initial line value of 250 Vrms, the
reference is set to 380 Vrms. In this case, the load is kept at its
half-load value (RL = 128 ).
Figs. 1115 show the three-phase stator voltages, switching
function, and stator voltage amplitude. Both simulation (see
Figs. 1113) and experimental tests (see Figs. 14 and 15)
exhibit good stator voltage regulation performance.

4894

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 58, NO. 10, OCTOBER 2011

Fig. 14. Experimental results: Three-phase stator voltages and switching


function for a voltage reference change from 250 to 380 Vrms with half load.
Fig. 11. Simulation results: Three-phase stator voltages and switching function for a voltage reference change from 250 to 380 Vrms with half load.

Fig. 15. Experimental results: Switching control policy and its filtered value
for a voltage reference change from 250 to 380 Vrms with half load.

Fig. 12. Simulation results: Stator voltage amplitude Vs for a voltage reference change from 250 to 380 Vrms with half load.

Fig. 13. Simulation results: Field voltage vF and equivalent control ueq for a
voltage reference change from 250 to 380 Vrms with half load.

The field voltage and the equivalent control (for simulation


results) and the switch driver signal and its filtered value (for the
experimental results) are shown in Figs. 13 and 15, respectively.
It is interesting to note that commutation (sliding mode) is lost
for a short period when the reference voltage changes, due to
the limited value of Vdc . However, it recovers fast, and the
equivalent control returns to the operation strip.
As appointed in [22], considering a dynamic load implies a
significant modification during the control design. The following test is performed in order to validate the control law beyond
the considered kind of load and to explore the robustness of the
obtained algorithm.
The reference line voltage is set to 380 Vrms. The initial load
is the half load value used in the previous tests, and suddenly, a
736-W induction motor (IM) is connected.
The three-phase stator voltages, switching function, and stator voltage amplitude are shown in Figs. 1620. It is worth
noticing again the good performance of the controlled stator
voltage amplitude.
The switching function oscillates around zero, and the consequent chattering phenomenon is just a little bit appreciated in
simulations due to the high value of the switching frequency.
However, when the load is inductive, the consequent chattering

MUOZ-AGUILAR et al.: SLIDING MODE CONTROL OF A STAND-ALONE WOUND ROTOR SYNCHRONOUS GENERATOR

4895

Fig. 19. Experimental results: Three-phase stator voltages and switching


function for a change from the one-half-load value to the IM connection.
Fig. 16. Simulation results: Three-phase stator voltages and switching function for a change from half-load value to the induction motor (IM) connection.

Fig. 20. Experimental results: Stator voltage amplitude and its reference and
switching control police and its filtered value for a change from the one-halfload value to the IM connection.

Fig. 17. Simulation results: Stator voltage amplitude Vs for a change from the
one-half-load value to the IM connection.

phenomenon diminishes because the inductive load acts as a


filter.
Figs. 18 and 20 contain the switching action and the equivalent control for simulations and experimental tests, respectively.
The sliding motion is lost while the stator voltage regulation is
not achieved. In the meantime, the actual control is saturated.
The reaching time can be reduced by increasing the Vdc voltage
value.
Then, the time to reach the reference depends on the Vdc
voltage value.

VI. C ONCLUSION

Fig. 18. Simulation results: Field voltage vF and equivalent control ueq for a
change from the one-half-load value to the IM connection.

This paper has presented a sliding mode controller designed


for a WRSM acting as a stand-alone generator. The obtained
control law regulates the stator voltage amplitude irrespective
of the load value, and the closed loop system is robust to machine parameters. Furthermore, only voltage and rotor position
measurements (to compute the dq-transformation) are required.
The control algorithm, which is extremely easy to implement
and does not require any gain tuning, is validated via simulations and experimental tests. The obtained results show a rapid

4896

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 58, NO. 10, OCTOBER 2011

response to load and voltage reference changes. The controller


was also tested for inductive loads, showing good performance.
A nice feature of the control algorithm, provided by the
use of the sign of the d-voltage component in the control
switching policy, is the ability to stabilize the system at the
two equilibrium points. This cannot be achieved with a simple
proportionalintegral controller where one of the two equilibrium points becomes unstable.
Closed loop dynamics, which becomes highly nonlinear, is
also discussed, and local stability is proved using small-signal
analysis.
Future research includes synthesis and analysis problems
caused by nonlinear and dynamical loads.
ACKNOWLEDGMENT

[17] H. Huerta, A. G. Loukianov, and J. M. Caedo, Multimachine powersystem control: Integral-SM approach, IEEE Trans. Ind. Electron.,
vol. 56, no. 6, pp. 22292236, Jun. 2009.
[18] G. Foo and M. F. Rahman, Sensorless sliding-mode MTPA control of
an IPM synchronous motor drive using a sliding-mode observer and HF
signal injection, IEEE Trans. Ind. Electron., vol. 57, no. 4, pp. 1270
1278, Apr. 2010.
[19] S. Benelghali, M. El Hachemi-Benbouzid, J. F. Charpentier,
T. Ahmed-Ali, and I. Munteanu, Experimental validation of a marine
current turbine simulator: Application to a permanent magnet
synchronous generator-based system second-order sliding mode control,
IEEE Trans. Ind. Electron., vol. 58, no. 1, pp. 118126, Jan. 2011.
[20] E. Mouni, S. Tnani, and G. Champenois, Improvement of synchronous
generators transient state by using voltage source and sliding mode control, in Proc. IEEE Int. Symp. Ind. Electron., Jun./Jul. 2008.
[21] IEEE. Ieee Guide: Test Procedures for Synchronous Machines, Std.1151995, 1995.
[22] R. S. Muoz-Aguilar, A. Dria-Cerezo, and E. Fossas, A sliding mode
control for a wound rotor synchronous generator with an isolated RL
load, in Proc. Amer. Control Conf., 2010, pp. 25512556.

The authors would like to thank Prof. V. I. Utkin for the


comments during the controller design and experimental steps
and E. Mir for helping in the implementation of the power
converter control stages.
R EFERENCES
[1] P. M. Anderson and A. A. Fouad, Power System Control and Stability.
Ames, IA: Iowa State Univ. Press, 1977.
[2] P. C. Krause, F. Nozari, T. L. Skvarenina, and D. W. Olive, The theory of
neglecting stator transients, IEEE Trans. Power App. Syst., vol. PAS-98,
no. 1, pp. 141148, Jan. 1979.
[3] W. Leonhard, Control of Electric Drives. Berlin, Germany: SpringerVerlag, 1995.
[4] F. Genduso, R. Miceli, C. Rando, and G. R. Galluzzo, Back EMF
sensorless-control algorithm for high-dynamic performance PMSM,
IEEE Trans. Ind. Electron., vol. 57, no. 6, pp. 20922100, Jun. 2010.
[5] L. Idkhajine, E. Monmasson, M.-W. Naouar, A. Prata, and K. Bouallaga,
Fully integrated FPGA-based controller for synchronous motor drive,
IEEE Trans. Ind. Electron., vol. 56, no. 10, pp. 40064017, Oct. 2009.
[6] V. D. Colli, R. Di Stefano, and F. Marignetti, A system-on-chip sensorless
control for a permanent-magnet synchronous motor, IEEE Trans. Ind.
Electron., vol. 57, no. 11, pp. 38223829, Nov. 2010.
[7] E. Ho and P. C. Sen, High-performance decoupling control techniques
for various rotating field machines, IEEE Trans. Ind. Electron., vol. 42,
no. 1, pp. 4049, Feb. 1995.
[8] A. Dria-Cerezo, C. Batlle, and G. Espinosa-Prez, Passivity-based control of a wound rotor synchronous motor, IET Control Theory Appl.,
vol. 4, no. 10, pp. 20492057, Oct. 2010.
[9] H. W. de Kock, A. J. Rix, and M. J. Kamper, Optimal torque control
of synchronous machines based on finite-element analysis, IEEE Trans.
Ind. Electron., vol. 57, no. 1, pp. 413419, Jan. 2010.
[10] Z. Sun, J. Wang, G. W. Jewell, and D. Howe, Enhanced optimal torque
control of fault-tolerant PM machine under flux-weakening operation,
IEEE Trans. Ind. Electron., vol. 57, no. 1, pp. 344353, Jan. 2010.
[11] M.-W. Naouar, E. Monmasson, A. Naassani, I. Slama-Belkhodja, and
N. Patin, FPGA-based current controllers for ac machine drivesA
review, IEEE Trans. Ind. Electron., vol. 54, no. 4, pp. 19071925,
Aug. 2007.
[12] J. Y. Hung, W. Gao, and J. C. Hung, Variable structure control: A survey,
IEEE Trans. Ind. Electron., vol. 40, no. 1, pp. 222, Feb. 1993.
[13] D. Biel and E. Fossas, SMC application in power electronics, in Variable Structure Systems: From Principles to Implementation, vol. 66,
A. Sabanovic, L. M. Fridman, and S. Spurgeon, Eds. London, U.K.: IEE
Control Series, 2004, pp. 265284.
[14] V. Utkin, J. Guldner, and J. Shi, Sliding Mode Control in Electromechanical Systems. New York: Taylor & Francis, 1999.
[15] A. E. Sanchez-Orta, J. De Len-Morales, and E. Lopez-Toledo, Discretetime nonlinear control scheme for small synchronous generator, Int. J.
Elect. Power Energy Syst., vol. 24, no. 9, pp. 751764, Nov. 2002.
[16] A. G. Loukianov, J. M. Caedo, L. M. Fridman, and A. Soto-Cota, Highorder block sliding-mode controller for a synchronous generator with an
exciter system, IEEE Trans. Ind. Electron., vol. 58, no. 1, pp. 337347,
Jan. 2011.

Ral Santiago Muoz-Aguilar received the B.Sc.


degree in electrical engineering and the M.Sc. degree
in industrial automation from the National University
of Colombia, Manizales, Colombia, in 2005 and
2007, respectively, and the M.Sc. degree in automation and robotics and the Ph.D. degree in advanced
automation and robotics from the Universitat Politcnica de Catalunya (UPC), Barcelona, Spain, both
in 2010.
In 2009, he joined the faculty of UPC, Terrassa,
Spain, where he is currently an Assistant Professor
with the Department of Electrical Engineering. His research interests include
nonlinear control systems, electrical drives, power electronics, electrical machines, and renewable energy systems among others.

Arnau Dria-Cerezo was born in Barcelona in


1974. He received the B.S. degree in electromechanical engineering and the Ph.D. degree from Universitat Politcnica de Catalunya (UPC) in 2001 and
2006, respectively. He received his DEA in Industrial
Automation from Laboratoire dAutomatisme Industrielle, INSA-Lyon, France, in 2001.
He is currently a Teaching Assistant with the Department of Electrical Engineering, Escola Politcnica Superior dEnginyeria de Vilanova i la Geltr,
UPC, Vilanova i la Geltr, Spain, and carries out his
research with the research group on Advanced Control of Energy Systems at
Institut dOrganizaci i Control de Sistemes Industrials (IOC), UPC. In 2003
and 2004, he was a Control Training Site-Research Fellow with Laboratoire
des Signaux et Systmes (L2S) at Suplec, France. In 2010, he was Visitor at
the Technische Universiteit Delft, The Netherlands. His main research interests
include control of electrical systems (related to its generation, conditioning,
management, and storage), modeling and advanced control of electrical machines and power converters, and modeling and control using passivity-based
techniques.

MUOZ-AGUILAR et al.: SLIDING MODE CONTROL OF A STAND-ALONE WOUND ROTOR SYNCHRONOUS GENERATOR

Enric Fossas (M08) received the M.S. and


Ph.D. degrees in mathematics from Universitat de
Barcelona, Barcelona, Spain, in 1981 and 1986,
respectively.
Since 1981, he taught mathematics in Universitat
de Barcelona and mathematics and automatic control in Universitat Politcnica de Catalunya (UPC),
where he is a Full Professor. He was the Head of
the Department of Applied Mathematics, UPC, from
September 1993 to August 1999 and the Director of
the Institute of Industrial and Control Engineering,
UPC, from July 2002 to July 2008. His research interests are in variable
structure systems and nonlinear control with applications to power electronics.
Dr. Fossas is a member of SIAM.

4897

Rafel Cardoner Papal was born in Barcelona,


Spain, in 1960. He received the M.Sc. degree in
technical telecomunications engineering from the
Universitat Ramon Llul, La Salle, Spain.
He is currently a Development Engineer with
the Institute of Industrial and Control Engineering,
Universitat Politcnica de Catalunya, Barcelona. His
research interests include digital control and power
electronics.

You might also like