Sie sind auf Seite 1von 10

IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO.

5, SEPTEMBER/OCTOBER 2010 1989


Sensorless Control of Induction Machines at Low
and Zero Speed by Using PWM Harmonics
for Rotor-Bar Slotting Detection
Reiko Raute, Member, IEEE, Cedric Caruana, Member, IEEE, Cyril Spiteri Staines, Member, IEEE,
Joseph Cilia, Mark Sumner, Senior Member, IEEE, and Greg M. Asher, Fellow, IEEE
AbstractThis paper presents the use of the inherent high-
frequency pulsewidth modulation (PWM) harmonics for sensor-
less control of ac machines. The amplitude and position of the
PWM voltage harmonics cannot be controlled independently and
are determined by the fundamental machine operation. However,
they do form a high-frequency excitation and can provide infor-
mation on saliencies within ac machines. This paper examines the
feasibility of sensorless control based on extracting the rotor-bar
slot position information for a cage induction machine using PWM
harmonics. The position-signal demodulation and compensation
schemes used are derived. Experimental results are provided for
an off-the-shelf induction motor operating under sensorless cur-
rent, speed, and position control, including zero excitation and
zero speed.
Index TermsInduction machine, pulsewidth modulation
(PWM), saliency, sensorless control, vector control.
I. INTRODUCTION
F
OR MANY years, the spatial impedance anisotropy in ac
machines has been considered for the determination of the
mechanical rotor or ux position. Published methods utilized
the machine response to injected test signals, including high-
frequency [1][10] and voltage-test-vector [11][13] injection.
Even though these methods were successful in implementing
true zero-speed sensorless control for ac-machine drives, the
use of additional test signals comes with disadvantages, such
as additional losses and acoustic noise, due to the test signal
creating additional disturbances in the machine currents.
Recent approaches [14][16] modied the pulsewidth mod-
ulation (PWM) scheme to embed the voltage test vectors
into the fundamental PWM cycles and avoid additional
Manuscript received October 2, 2009; revised December 17, 2009; accepted
January 26, 2010. Date of publication July 12, 2010; date of current version
September 17, 2010. Paper 2009-IDC-327.R1, presented at the 2008 Industry
Applications Society Annual Meeting, Edmonton, AB, Canada, October 59,
and approved for publication in the IEEE TRANSACTIONS ON INDUSTRY
APPLICATIONS by the Industrial Drives Committee of the IEEE Industry
Applications Society. This work was supported by the European Commission
under a Marie Curie Research Training Network for realizing this work under
the MEST-CT-2004-504243 Research Project.
R. Raute, C. Caruana, C. S. Staines, and J. Cilia are with the Faculty of
Electrical Engineering, University of Malta, MSD 06 Msida, Malta (e-mail:
cccaru@eng.um.edu.mt).
M. Sumner and G. M. Asher are with the School of Electrical Engi-
neering, University of Nottingham, Nottingham, NG7 2RD, U.K. (e-mail:
Mark.Sumner@nottingham.ac.uk; greg.asher@nottingham.ac.uk).
Color versions of one or more of the gures in this paper are available online
at http://ieeexplore.ieee.org.
Digital Object Identier 10.1109/TIA.2010.2057495
switching. However, signicant current distortions were still
observed [15].
This paper presents a novel technique that exploits the high-
frequency signals that are inherently imposed on the machine
due to the nature of the inverters modulated output voltage.
The high-frequency voltage harmonics, at multiples of the
PWM switching frequency, provide sufcient pulsating exci-
tation for the machine to produce a current response with
position-dependent modulation [17]. This paper will examine
the feasibility of extracting rotor-position information from the
resulting high-frequency components for the sensorless control
of ac machines.
Contrary to normal signal-injection techniques, the high-
frequency voltage signal amplitude and position cannot be
independently set [18]. The second PWM harmonic was identi-
ed as the strongest pulsating HF signal [17] and is therefore
utilized as the HF excitation signal for position detection.
However, the same technique can be applied to the other
harmonics. The nature of the selected PWM harmonic as an
excitation signal is described in Section II, and the algorithm
for using the test signal to derive a rotor-position signal is
presented in Section III. Sections IVVIII discuss the prac-
tical implementation of the algorithm, including the various
signal processing and control algorithms required to extract the
desired signal from the raw signal which contains distorting
effects, such as higher order harmonic saturation harmonics. It
should be emphasized that this paper describes the application
of the new sensorless-control technique to the sensorless vector
control of unmodied induction motors (IMs), i.e., motors
which have a small degree of rotor slotting saliency. The test
motor used in the experimental system has a skewed rotor
and has semiclosed slots and therefore presents an extremely
demanding control challenge. Experimental results presented
in Section IX demonstrate that, despite the complexity of the
proposed control approach, good sensorless performance can
be achieved in the difcult low- and zero-speed regions.
II. BEHAVIOR OF PWM HARMONICS
In standard voltage-source inverters (VSIs), the principle of
PWM is used to generate the fundamental excitation voltage.
The inverter output is switched between the positive and nega-
tive potential of the dc-link voltage at a high carrier frequency.
The desired fundamental output voltage is the average of the
PWM periods. Fig. 1 shows the inverter output voltage for one
0093-9994/$26.00 2010 IEEE
1990 IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO. 5, SEPTEMBER/OCTOBER 2010
Fig. 1. One PWM cycle of the generated inverter output voltage: (a) phase
output voltages and (b) , components of resulting voltage vector.
PWM period. The PWM inverter voltage generation produces
a switched voltage pattern with sharp edges. This generated
voltage signal has frequency components that is theoretically
up to innity. Since the PWM period width is xed and
can be regarded as repeated pattern, the Fourier series can
be calculated to investigate the frequency components. As a
basic rule of repeated patterns, only frequency components of
multiples of the pattern frequency will be present in the Fourier
series. That means that the complete inverter output voltage can
be fragmented into frequency components of multiples of the
PWM switching frequency plus a dc component.
As a basic approach, the inverter phase-switching duty cycle
d
x
can be calculated by (1), where the modulation index m
x
depends directly on the desired phase-voltage amplitude (2).
The subscript x refers to phases A, B, and C
d
x
=0.5 + 0.5m
x
(1)
m
x
=
2v
x
V
DC
. (2)
According to the general form of a Fourier series, the PWM
output voltage per phase v
x
can be described as
v
x
(t) =
a
PWM0x
2
+

n=1
(a
PWMnx
cos(n
PWM
t)
+ b
PWMnx
sin(n
PWM
t)) . (3)
The desired fundamental inverter output voltage is the PWM
period dc component a
PWM0x
/2 which is (d
x
0.5)V
DC
.
Since the signal is regarded as an even function, all b
PWMnx
co-
efcients are zero. The amplitude of each nth PWM harmonic
can be calculated by the coefcient a
PWMnx
a
PWMnx
=
2V
DC
n
sin (n(0.5 + 0.5m
x
)) . (4)
Fig. 2. Pulsating v
PWM2
vector rotating with fundamental voltage vector v
in frame.
In the space-vector approach, the fundamental three-phase
voltages can be combined to a complex phasor in the stator
xed frame as follows:
v

=
2
3
V
DC
2

m
A
+ m
B
e
j
2
3

+ m
C
e
j
4
3

. (5)
Thus, the three-phase PWM carrier harmonic components
can be also combined to the complex phasors in the ref-
erence frame to
a
PWMn
=
2
3

a
PWMnA
+a
PWMnB
e
j
2
3

+a
PWMnC
e
j
4
3

.
(6)
It was found that, at low speed, the second PWM harmonic
has the largest amplitude. Looking back to Fig. 1 helps in
understanding this conclusion since it can be seen that, in each
PWM period, the active voltage vectors occur twice, centered
from the middle of the PWM cycle. Therefore, only the sec-
ond PWM harmonic is used in the sensorless algorithm. All
variables referred to this PWM harmonic are further denoted as
PWM2. From (6) and (4), the second PWM carrier harmonic
can be described as a pulsating vector, rotating approximately
synchronously with the fundamental voltage vector in the stator
xed frame, as shown in
v
PWM2
=
2
3
V
DC

sin(m
A
) + sin(m
B
)e
j
2
3

+ sin(m
C
)e
j
4
3

cos(2
PWM
t). (7)
Fig. 2 shows the fundamental voltage vector v (5) and the
resulting pulsating HF vector v
PWM2
(7).
III. USE OF PWM HARMONICS FOR
SALIENCY DETECTION
The second PWM harmonic can be regarded as an injected
HF pulsating vector. The resulting current PWM2 carrier
harmonic, together with the injected HF voltage vector, can
be used for detecting the impedance inside the machine, sim-
ilar to HF pulsating-injection drives [5][8]. However, as the
HF pulsating-vector amplitude and phase are now determined
RAUTE et al.: SENSORLESS CONTROL OF INDUCTION MACHINES BY USING PWM HARMONICS 1991
Fig. 3. Block diagram of PWM2 signal demodulation.
by the fundamental operation, a novel position observer is
required.
Fig. 3 shows a block diagram of the implemented signal
processing. The measured three-phase voltages and currents are
directly combined in the time domain to the resulting stator
voltage and current vector (v

, i

). The components of
the vectors are bandpass ltered. The lters center frequency
is set to twice the PWM switching frequency. The width of the
passband needs to be carefully selected. The bandpass lter
removes unwanted frequency components and reduces noise,
but it also determines the maximum frequency of the detectable
saliency signal modulation. The bandpass-ltered signals are
further demodulated by a heterodyning technique. The resulting
f
PWM2
voltage and current vector can be described by (8) and
(9), where 2v

PWM2
and 2i

PWM2
are the amplitudes of the
pulsating signals
v
PWM2
=2v

PWM2
cos(2f
PWM2
t +
vPWM2
) (8)
i
PWM2
=2i

PWM2
cos(2f
PWM2
t +
iPWM2
). (9)
The high-frequency voltage vector is directly multiplied
by a cosine reference signal cos(2f
PWM2
t +
vPWM2
), and
the high-frequency current vector signal is multiplied by
cos(2f
PWM2
t +
iPWM2
). The cosine reference signals are
stored in a lookup table (LUT) for the known signal samples.
The phases of the cosine multiplication signals were set on ex-
perimental data and ne-tuned manually. The resulting voltage
and current signals are therefore given by
v
PWM2
cos(2f
PWM2
t +
vPWM2
)
= v

PWM2
+ v

PWM2
cos(22f
PWM2
t + 2
vPWM2
) (10)
i
PWM2
cos(2f
PWM2
t +
iPWM2
)
= i

PWM2
+ i

PWM2
cos(22f
PWM2
t + 2
iPWM2
). (11)
The HF carrier frequency component, which converts to
2f
PWM2
, is removed by a discrete average lter. The av-
erage lter calculates the average of a time period T that
includes exactly an integer number of periods of the HF signal
cos(22f
PWM2
t). As a result, only the amplitude modulation
signals v

PWM2
and i

PWM2
of the f
PWM2
voltage and current
signals are derived, as shown in the following:
v

PWM2
=
1
T
T

0
(v

PWM2
+v

PWM2
cos(22f
PWM2
t+2
vPWM2
)) dt (12)
TABLE I
PARAMETERS OF THE EMPLOYED INDUCTION MACHINE
i

PWM2
=
1
T
T

0
(i

PWM2
+i

PWM2
cos(22f
PWM2
t+2
iPWM2
)) dt. (13)
An equivalent-impedance vector z

PWM2
can be dened
based on the demodulated voltage and current PWM carrier
harmonic vectors v

PWM2
and i

PWM2
(14). The ratio between
v

PWM2
and i

PWM2
takes the uncontrolled variation in the
injected HF excitation into account. The complex vector
division can be calculated in any reference frame. The result
will be only affected by the phase and magnitude difference
of the two vectors. The nal HF equivalent-impedance vector
z

PWM2
is calculated by the complex vector division directly in
the frame, i.e.,
z

PWM2
=
v

PWM2
i

PWM2
. (14)
IV. PRACTICAL IMPLEMENTATION
The parameters of the off-the-shelf induction machine are
stated in Table I. The rotor is a skewed squirrel-cage design
with semiopened rotor slots. A controlled dc drive is coupled to
the induction machine under test as the load.
For the control unit, a Texas Instruments DSK 6713C board,
in conjunction with a customized eld-programmable gate
array (FPGA) board, is used. The FPGA board provides pe-
ripheral interfaces that allow using the DSK 6713C board for
the control of electric drives. It contains a standard PWM
signal-generation unit for the control of a three-phase two-
level VSI and 16-b analog-to-digital converters (ADCs) for
the current and voltage measurements. The analog current-
and voltage-measurement signals are ltered by a fourth-order
10-kHz low-pass antialiasing lters before the digital sampling.
The generated PWM switching signals are directly fed into a
commercial inverter. No dead-time compensation strategy is
implemented at this stage. The PWM switching frequency is
set to 1.818 kHz. This is a rather low value for the 5.5-kW
machine used. This value has been chosen to ensure that the
investigation of the basic algorithm and its associated signal-
processing and compensation strategies is not restricted by the
processing capabilities of the DSP and FPGA platform used.
This low PWM frequency allows the voltage to be sampled
at a signicantly higher rate such that a reasonably accurate
value for the second PWM harmonic can be obtained online.
It is recognized that, if a more realistic PWM frequency were
considered (e.g., 10 kHz), then the signal-to-noise ratio may be
poorer for the current measurements, but, at present, this is con-
sidered to be a feasibility study rather that a full commercializa-
tion project. The digital bandpass lter used is a fourth-order
1992 IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO. 5, SEPTEMBER/OCTOBER 2010
Fig. 4. Visualization of rotor-bar equivalent-impedance modulation and re-
sulting v

PWM2
and i

PWM2
for steady-state operation.
Butterworth design tuned to a center frequency of 3.636 kHz
with 400-Hz bandwidth. All currents and voltages are sampled
at 30 kHz. The ADC sampling frequency is exactly 16.5 times
the PWM switching frequency. The sensorless algorithm is
executed every second PWM cycle. Therefore, the HF signal
demodulation is performed over 33 signal samples, and the
cycle frequency of the sensorless algorithm is 909.09 Hz.
V. MEASURED ROTOR-BAR MODULATION EFFECTS
The focus of this paper is to detect the asynchronous mod-
ulation due to the conductor bars embedded in the rotor iron
package of the machine. It is assumed that the rotor bars cause
a circular equivalent-impedance modulation with an amplitude
Z

RB
. The angle
RB
is the rotor-bar position within one rotor-
bar period, which is the distance between two adjacent rotor
bars. The resulting voltage equation systemfor the demodulated
PWM2 variables is shown by

PWM2

PWM2

RB
cos(
RB
) Z

RB
sin(
RB
)
Z

RB
sin(
RB
) Z

+ Z

RB
cos(
RB
)

PWM2
i

PWM2

. (15)
Fig. 4 shows the equivalent-impedance modulation. The
PWM2 voltage and current vectors (v

PWM2
and i

PWM2
) are
shown for the condition that v

PWM2
is 45

ele
in the stator
frame. The injected pulsating HF voltage vector v

PWM2
will rotate with the fundamental frequency. In steady state, the
amplitude of this vector will be approximately constant, and the
locus will be almost circular in the frame (dashed line).
The resulting pulsating HF current vector i

PWM2
will follow
the imposed voltage vector. However, due to impedance varia-
tion, the current-vector locus will be slightly different, as shown
by the oval shape (dotted line). The difference between the HF
voltage and current vector is proportional to the equivalent-
impedance modulation Z

RB
. This modulation is shown in the
drawings for different v

PWM2
positions in the frame. The
oval i

PWM2
locus will rotate by 180

ele
per rotor-bar period. As
can be seen from the gure, a further rotation occurs in the
detected equivalent-impedance modulation depending on the
PWM2 vector positions in the stator frame. For the explana-
tion of this additional modulation, the xy reference frame, also
shown in Fig. 4, is introduced. The equivalent-impedance tensor
Z

PWM2
can then be transformed to Z

PWM2xy
by applying
(17), including the rotation matrix (16), which are both shown
at the bottom of the page.
Equation (18) shows the voltage equation for the demod-
ulated PWM2 quantities in the xy reference frame shown
in Fig. 4

PWM2x
v

PWM2y

= Z

PWM2xy

|i

PWM2
|
0

. (18)
The result of the complex vector division (14) is stated by
(19), which shows that an equivalent-impedance vector with
an offset Z

and a circular modulation with the radius Z

RB
rotating backwards with
RB
+ 2i

PWM2
occurs, i.e.,
z

PWM2
=
v

PWM2
+ jv

PWM2
i

PWM2
+ ji

PWM2
=
v

PWM2x
+ jv

PWM2y
|i

PWM2
|
=Z

RB
cos

RB
+ 2i

PWM2

+ jZ

RB
sin

RB
+ 2i

PWM2

. (19)
After compensating for the offset, the additional 2i

PWM2
phase modulation can be easily removed since the HF current
vector position i

PWM2
is directly known.
B
xy
=

cos

PWM2

sin

PWM2

sin

PWM2

cos

PWM2

(16)
Z

PWM2xy
=B
xy
Z

PWM2
B
T
xy
=

RB
cos

RB
+ 2i

PWM2

RB
sin

RB
+ 2i

PWM2

RB
sin

RB
+ 2i

PWM2

+ Z

RB
cos

RB
+ 2i

PWM2

(17)
RAUTE et al.: SENSORLESS CONTROL OF INDUCTION MACHINES BY USING PWM HARMONICS 1993
Fig. 5. Modeled saturation HF modulation.
VI. SATURATION-MODULATION EFFECTS
In standard induction machines, strong magnetic-saturation
effects are present. For simplicity, it can be assumed that the
saturation effect on the impedance modulation is superim-
posed to the rotor-bar modulation. The saturation equivalent-
impedance modulation is theoretically also modeled as an
ellipse, rotating with the fundamental current vector in the
frame. As the HF voltage vector is also rotating with the
fundamental frequency, it can be assumed that the resulting
saturation effect of the z

PWM2
equivalent-impedance vector is
static. Therefore, in steady state, the magnetic saturation will
only affect the dc component of the vector z

PWM2
. Fig. 5 shows
the effect of the magnetic saturation inuence to the z

PWM2
measurement. The axis of the saturation shape is dened by the
maximal magnetic saturation.
In steady state, the position of the HF equivalent-impedance
measurement frame (i

PWM2
) with respect to the saturation
axis S is constant. However, at different speeds, the relative
positions of S and i

PWM2
change. This is mostly due to
the speed-dependent back electromotive force. Therefore, the
detected z

PWM2
saturation inuence will not be constant under
dynamic operation. The total magnetic saturation within the
machine will have a further effect on the diameter of the
saturation-saliency shape and, therefore, the offset Z

.
VII. DECOUPLING OF ROTOR-BAR MODULATION
In the theoretical explanation, it is assumed that the average
machine equivalent impedance Z

and the modulation effects


are superimposed as stated in
z

PWM2
= Z

+ z

PWM2Sat
+ z

PWM2RB
+ z

PWM2Inv
.
(20)
In the actual machine used, the measured rotor-bar modu-
lation Z

RB
is very low (1% to 1.5% of the offset value Z

)
due to the rotor construction and, particularly, the skewing.
The saturation modulation z

PWM2Sat
is up to 4% of Z

. It
was observed that further modulation occurs with each current
commutation sector. This effect is assumed to be due to inverter
nonlinearity, caused by the dead time during the switching
between the two insulated-gate bipolar transistors per-phase
leg and current-clamping effects [2]. The amplitude of this
inverter modulation z

PWM2 Inv
is in the range of about 2%
to 3% of Z

and is actually higher than the slotting effect to be


tracked. Therefore, the decoupling of the desired signal for the
rotor-position estimation from the other distorting effects is a
critical task. In this paper, a basic LUT compensation scheme
is implemented to extract only the desired rotor-bar modulation.
As explained in Section VI, the saturation modulation depends
on the imposed stator currents and the relative position i

PWM2
to the saturation axis position S, as shown in Fig. 5. In
the regarded operation, i
d
is kept constant, and thus, the total
machine current amplitude depends directly on the torque-
producing current component i
q
. Since the current controller
has a fast dynamic response, i

q
is used as a reference for the
LUT for noise reduction. A further compensation dimension
is i

PWM2
S. The position of the saturation axis S is not
known, but it is assumed that it keeps a constant relationship
with the stator current vector i for each particular value of
i
q
. Therefore, i

PWM2
i is used as a second reference for
the LUT. For the compensation of the inverter nonlinearity
effect, i

is used as the third LUT dimension. The nal


implemented LUT contains the values of z

PWM2 LUT
(21)
addressed by the three reference dimensions
z

PWM2LUT
(i
q
, i

, i

PWM2
i)
= Z

+ z

PWM2Sat
+ z

PWM2Inv
. (21)
For this feasibility study, the LUT was created using a special
commissioning test. The drive is operated in fully sensored
vector control mode over the whole of the required operating
range required for the proposed sensorless algorithm i.e., vari-
able speed and i
q
. The i
q
current, the stator current angle i

,
and the angle i

PWM2
i are acquired together with the
measured equivalent-impedance vector z

PWM2
. Therefore, the
z

PWM2
modulation is measured, including the rotor bar, satu-
ration, and inverter nonlinearity modulation for the entire oper-
ating range. The measured data are processed ofine according
to the measured i
q
, i

, and i

PWM2
i parameters. Data
points with the same i
q
, i

, and i

PWM2
i references
are averaged. Since the rotor-bar modulation is asynchronous to
the measured i
q
, i

, and i

PWM2
i parameters and the
data were collected over a nite time duration, it is assumed that
the rotor-bar modulation is removed due to averaging of several
samples per i
q
, i

, and i

PWM2
i reference. Thus, a
z

PWM2 LUT
modulation prole is generated, which does not in-
clude the rotor-bar modulation. Figs. 6 and 7 show the real and
imaginary parts of the measured z

PWM2 LUT
prole dependent
on i

and i

PWM2
i for the rated current. The parameter
range is discretized by 2

ele
for both reference parameters. The
at sections in the gures show conditions that did not occur
during operation. The i

PWM2
i range is limited due to
the fundamental operation of the machine, and therefore, the
saliency function is only detected in a small section of the
360

ele
range. Also, during sensorless operation, only points in
the detected section can occur, and thus, the knowledge of this
compensation prole is sufcient.
1994 IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO. 5, SEPTEMBER/OCTOBER 2010
Fig. 6. Real part of z

PWM2 LUT
(i
q
= 12.5 A, i

, i

PWM2
i).
Fig. 7. Imaginary part of z

PWM2 LUT
(i
q
= 12.5 A, i

, i

PWM2

i).
Fig. 8. Real part of z

PWM2 LUT
(i
q
, i

= 0

ele
, i

PWM2
i).
In the shown equivalent-impedance prole of z

PWM2 LUT
,
one can notice the parameter dependence of i

and
i

PWM2
i. It can be seen that the real-part offset Z

is about
190 V/A for the given i
q
. Also, the modulation pattern of the
current commutation sectors depends on i

, as can be seen.
The magnetic saturation-saliency modulation z

PWM2 Sat
has
the expected shape. From these gures, it can be seen that
z

PWM2 Sat
is up to 10 V/A, and the saturation axis, where
Re(z

PWM2 LUT
) is minimum and Im(z

PWM2 LUT
) is
zero, is approximately at i

PWM2
i = 0

ele
. This indicates
that the magnetic saturation axis is near the stator current vector.
Fig. 9. Imaginary part of z

PWM2 LUT
(i
q
, i

= 0

ele
, i

PWM2
i).
Fig. 10. Signal tracking PLLs in the sensorless algorithm.
Fig. 11. Block diagram of PLL1 structure.
Figs. 8 and 9 show the equivalent-impedance compensation
values as function of i
q
and i

PWM2
i. The 3-D graphs
show the data for the fundamental stator current position
i

= 0

ele
. In Fig. 8, it can be seen that the offset in the
real part of z

PWM2
also changes with i
q
. Therefore, it can be
assumed that the offset Z

is also a function of the machine-


current amplitude.
VIII. ROTOR-POSITION RECONSTRUCTION
Disturbances were observed in the LUT decoupling. The
disturbances are usually only of small duration. However, a
corrupted z

PWM2 RB
decoupling may cause rotor-bar skip-
ping, which produces a position-estimation step of one rotor-bar
period (22.5

ele
for the machine used). To reduce the likelihood
of rotor-bar skipping, phase-locked loops (PLLs) are used in
the sensorless rotor-position estimator, as shown in Fig. 10.
One PLL (PLL1) is used to track and lter the measured
z

PWM2 RB
modulation, which contains the rotor-bar modu-
lation signal (
RB
+ 2i

PWM2
). A second PLL (PLL2) is
used to condition the nal derived rotor-bar position signal. The
second PLL block is also used to obtain the mechanical position
and a ltered-speed estimate.
Figs. 11 and 12 show the two PLL structures in detail. For
the rst PLL, the real and imaginary parts of the decoupled
RAUTE et al.: SENSORLESS CONTROL OF INDUCTION MACHINES BY USING PWM HARMONICS 1995
Fig. 12. Block diagram of PLL2 structure.
z

PWM2 RB
are directly used as input signals. The out-
put is the position of the z

PWM2 RB
modulation (
RB
+
2i

PWM2
). It was observed that, under dynamic operation,
the LUT compensation does not provide a perfect signal decou-
pling in the real part of z

PWM2
. Even if the offset compensation
error is always small (< 3% of Z

), it can reach absolute


values larger than the rotor-bar modulation amplitude Z

RB
.
As a result, the quadrature PLL1 error signals derived from
the real and imaginary parts of z

PWM2 RB
are conditioned
by the scaling factors k
Re
and k
Im
. k
Im
is set to a larger
value than k
Re
. Therefore, the combined error signal to the
proportionalintegral controller is mainly derived from the
imaginary part of z

PWM2 RB
. The tradeoff is that, even under
correct lock in of the PLL, an error signal at twice the input
frequency can be seen.
The second PLL (PLL2) is applied after the 2i

PWM2
phase modulation has been removed. This PLL serves only to
further lter the estimated rotor-bar position signal. This re-
duces noise that might be introduced by the 2i

PWM2
phase
removal and further smoothes the effect of the incurred ripple
in PLL1. The tracking signal of PLL2 is directly converted
into a mechanical speed estimate and integrated into the nal
estimated rotor position
e
R
. Furthermore, a low-pass-ltered
speed signal
e
R F
is provided.
The PLL system for the signal conditioning provides a
signicant improvement of the rotor-bar position-signal track-
ing. It reduces the occurrences of rotor-bar skipping and
smoothes noise. This signal-conditioning structure could be
even more enhanced, for instance, by integrating a mechanical
observer [19].
IX. EXPERIMENTAL RESULTS
A. Current Control
Figs. 13 and 14 show the results when the sensorless IM
drive is operating in sensorless current-control mode. The speed
is controlled by the coupled dc machine. Fig. 13 shows an
experiment to test the dynamic response of the sensorless drive.
The ux-producing current component i

d
is kept constant, and
i

q
is changed from 0 to 12.5 A (0% to 100%). The i

q
transient
is limited to 100% per second. The mechanical speed is set to
52 rev/min by the coupled dc machine. Since at rated i
q
current, the slip frequency is about 1.74 Hz, the electrical
frequency is approximately zero when the maximum load is
applied to the machine.
Fig. 13. Sensorless drive operating in current control at constant speed and
varying load. (a) Measured mechanical rotor position
R
and estimated rotor
position
e
R
. (b) Position-estimation error (
e
R

R
). (c) i
e
dq
current com-
ponents of the sensorless vector-controlled drive. (d) Measured stator current
vector i

components. (e) and (f) Real and imaginary components of the


complex equivalent-impedance vector z

PWM2
and the LUT compensation
components.
It can be seen that at full applied i

q
, the stator currents are
almost dc. Plots (e) and (f) show that the LUT compen-
sation values (z

PWM2 LUT
) follow the oating mean of
the equivalent-impedance vector z

PWM2
. However, during the
transitions of i
e
q
, the compensation values do not follow the
z

PWM2
average precisely. Therefore, the correct rotor-bar
equivalent-impedance modulation z

PWM2 RB
decoupling is
not always ensured. The LUT for the modulation decou-
pling was generated using several experiments with constant
i
q
. It was observed that, under i
q
transients, the equivalent-
impedance modulation behaves differently than under steady-
state conditions. It is assumed that the saturation effect of
z

PWM2
does not change instantly with i
dq
transitions. As can
be seen, the LUT compensation error is only pronounced in
the real part of z

PWM2
. However, the PLL signal conditioning
ensures that the tracking of the correct rotor-bar position can be
achieved even under the i
q
transient conditions. It was found
that the error of the z

PWM2
compensation is dependent on the
step amplitude and rate of change in i
q
.
Fig. 14 shows the sensorless drive operating in current-
control mode during speed transients. The current controller
references i

d
and i

q
are both set to rated values. The speed
is changed from zero to 300 rev/min, 300 rev/min, and zero
speed again. The speed transients are limited to 600 rev/min/s. It
1996 IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO. 5, SEPTEMBER/OCTOBER 2010
Fig. 14. Sensorless drive operating at rated current and varying speed.
(a) Measured mechanical rotor position
R
and estimated rotor position
e
R
.
(b) Position-estimation error (
e
R

R
). (c) i
e
dq
current components of the
sensorless-vector-controlled drive. (d) Measured and estimated rotor speed.
(e) and (f) Real and imaginary components of the complex equivalent-
impedance vector z

PWM2
and the LUT compensation components.
can be seen that the drive operates stably at all times, including
at zero mechanical speed. From plots (e) and (f), it can be seen
that the LUT compensation follows the correct z

PWM2
values
even during the speed changes.
B. Speed Control
Fig. 15 shows the results of the implemented sensorless
IM drive operating in speed-control mode. The coupled dc
machine is set to provide a constant electrical-load torque. The
sensorless IM drive is controlling the speed by an implemented
speed loop providing the reference i

q
. The estimated speed

e
R F
provided by PLL2 is used for speed feedback. The speed
reference is changed between 60 rev/min. The transitions of
the reference speed are limited to 120 rev/min/s. The speed
controller bandwidth is set to a low value due to the limited
i
q
dynamic caused by the transient effect of the z

PWM2 RB
decoupling. It can be seen that the reference speed is obtained
by the sensorless drive. As can be seen, the drive provides up to
100% of rated current of the induction machine (i
q
= 12.5 A)
in this experiment. The visible stiction at zero speed is due
to friction in the mechanical coupling between the test and
load machine of the used rig. The same effect occurred also
in sensored operation and is therefore not associated to the
sensorless algorithm.
Fig. 15. Sensorless drive in speed control. (a) Measured mechanical rotor
position
R
and estimated rotor position
e
R
. (b) Position-estimation error
(
e
R

R
). (c) i
e
dq
current components of the sensorless vector-controlled
drive. (d) Measured and estimated mechanical speed n
R
and n
e
R
.
Fig. 16. Sensorless drive in position control. (a) Measured mechanical rotor
position
R
and estimated rotor position
e
R
. (b) Position-estimation error
(
e
R

R
). (c) i
e
dq
current components of the sensorless vector-controlled
drive. (d) Measured and estimated mechanical speed n
R
and n
e
R
.
C. Position Control
Fig. 16 shows the implemented sensorless drive operating
in position-control mode. The coupled dc machine is set to
provide a constant electrical-load torque. The position con-
troller of the sensorless IM drive provides a reference for
the speed-control loop. The reference position is changed by
one mechanical revolution (two electrical periods) forward and
backward. The reference position ramp is limited to 360

ele
per
second. The position-control loop contains only a proportional
controller since the plant transfer function provides an integral
RAUTE et al.: SENSORLESS CONTROL OF INDUCTION MACHINES BY USING PWM HARMONICS 1997
part. It can be seen that in this experiment, the sensorless drive
also provides up to rated i
q
.
X. CONCLUSION
This paper has proposed a new technique for accurate sen-
sorless control of a standard IM at zero and low speeds. The
technique detects the rotor slotting saliency but does not require
any additional test-signal injection or PWM pattern modica-
tion. The pulsating high-frequency excitation due to the second
harmonic of the PWM voltage waveform is used to detect
the equivalent impedance, which is calculated by a complex
division using the demodulated high-frequency current and
voltage vector variables. A precalibrated LUT is required to
decouple the rotor-position saliency from an offset and other
saliencies due to iron saturation and other distorting effects
such as inverter nonlinearities. A dual PLL structure is used to
improve the signal tracking.
Experimental results for a skewed off-the-shelf IM with
semiclosed slots show that even though the rotor-bar mod-
ulation is very small, sufcient position information can be
retrieved for position and speed control at zero and low speed.
This paper has demonstrated the feasibility of the proposed
algorithm, although a low switching frequency has been em-
ployed and detailed precommissioning is required. It is an-
ticipated that with the continuing advances in microprocessor
technology, the data sampling and processing algorithms may
be implemented at higher frequencies to allow the switching
frequency to increase to values on the order of 10 kHz. The
subject of precommissioning, online self-tuning, and the use of
alternative compensation devices such as expert systems, con-
tinues to be a major research area for most types of sensorless
drive control based on saliency tracking.
REFERENCES
[1] C. Caruana, G. M. Asher, and M. Sumner, Performance of HF signal in-
jection techniques for zero-low-frequency vector control of induction ma-
chines under sensorless conditions, IEEE Trans. Ind. Electron., vol. 53,
no. 1, pp. 225238, Feb. 2006.
[2] N. Teske, G. M. Asher, M. Sumner, and K. J. Bradley, Analysis and
suppression of high-frequency inverter modulation in sensorless position-
controlled induction machine drives, IEEE Trans. Ind. Appl., vol. 39,
no. 1, pp. 1018, Jan./Feb. 2003.
[3] Q. Gao, G. Asher, and M. Sumner, Sensorless position and speed control
of induction motors using high-frequency injection and without ofine
precommissioning, IEEE Trans. Ind. Electron., vol. 54, no. 5, pp. 2474
2481, Oct. 2007.
[4] M. W. Degner and R. D. Lorenz, Position estimation in induction
machines utilizing rotor bar slot harmonics and carrier-frequency sig-
nal injection, IEEE Trans. Ind. Appl., vol. 36, no. 3, pp. 736742,
May/Jun. 2000.
[5] M. J. Corley and R. D. Lorenz, Rotor position and velocity estimation
for a salient-pole permanent magnet synchronous machine at standstill
and high speeds, IEEE Trans. Ind. Appl., vol. 34, no. 4, pp. 784789,
Jul./Aug. 1998.
[6] J.-H. Jang, S.-K. Sul, J.-I. Ha, K. Ide, and M. Sawamura, Sensorless drive
of surface-mounted permanent-magnet motor by high-frequency signal
injection based on magnetic saliency, IEEE Trans. Ind. Appl., vol. 39,
no. 4, pp. 10311039, Jul./Aug. 2003.
[7] M. Linke, R. Kennel, and J. Holtz, Sensorless speed and position control
of synchronous machines using alternating carrier injection, in Proc.
IEMDC, Madison, WI, 2003, pp. 12111217.
[8] J. Holtz, Acquisition of position error and magnet polarity for sensorless
control of PM synchronous machines, IEEE Trans. Ind. Appl., vol. 44,
no. 4, pp. 11721180, Jul./Aug. 2008.
[9] A. Consoli, G. Scarcella, and A. Testa, Industry application of zero-speed
sensorless control techniques for PM synchronous motors, IEEE Trans.
Ind. Appl., vol. 37, no. 2, pp. 513521, Mar./Apr. 2001.
[10] F. Briz, W. Degner, P. Garca, and J. M. Guerrero, Rotor position esti-
mation of AC machines using the zero-sequence carrier-signal voltage,
IEEE Trans. Ind. Appl., vol. 41, no. 6, pp. 16371646, Nov./Dec. 2005.
[11] M. Schroedl, Sensorless control of AC machines at low speed and stand-
still based on the INFORM method, in Conf. Rec. IEEE IAS Annu.
Meeting, San Diego, CA, Oct. 610, 1996, pp. 270277.
[12] J. Holtz, Sensorless position control of induction motorsAn emerging
technology, IEEE Trans. Ind. Electron., vol. 45, no. 6, pp. 840852,
Dec. 1998.
[13] C. Caruana, G. M. Asher, and J. C. Clare, Sensorless ux position
estimation at low and zero frequency by measuring zero-sequence current
in delta-connected cage induction machines, IEEE Trans. Ind. Appl.,
vol. 41, no. 2, pp. 609617, Mar./Apr. 2005.
[14] J. Holtz and J. Juliet, Sensorless acquisition of the rotor position angle of
induction motors with arbitrary stator windings, IEEE Trans. Ind. Appl.,
vol. 41, no. 6, pp. 16751682, Nov./Dec. 2005.
[15] Q. Gao, G. M. Asher, M. Sumner, and P. Makys, Position estimation
of AC machines over a wide frequency range based on space vector
PWM excitation, IEEE Trans. Ind. Appl., vol. 43, no. 4, pp. 10011011,
Jul./Aug. 2007.
[16] Y. Hua, G. M. Asher, M. Sumner, and Q. Gao, Sensorless control of
surface mounted permanent magnetic machine using the standard space
vector PWN, in Conf. Rec. IEEE IAS Annu. Meeting, New Orleans, LA,
Sep. 2327, 2007, pp. 661667.
[17] R. Raute, C. Caruana, J. Cilia, C. S. Staines, M. Sumner, and G. Asher,
Impedance saliency detection for sensorless control of AC machines
utilizing existing PWM switching harmonics, in Proc. PEMD, York,
U.K., Apr. 24, 2008, pp. 557561.
[18] R. Raute, C. Caruana, C. S. Staines, J. Cilia, M. Sumner, and G. Asher,
Sensorless control of induction machines by using PWM harmonics for
rotor bar slotting detection, in Conf. Rec. IEEE IAS Annu. Meeting,
Edmonton, AB, Canada, Oct. 59, 2008, [CD-ROM].
[19] P. L. Jansen and R. D. Lorenz, Transducerless position and velocity
estimation in induction and salient AC machines, IEEE Trans. Ind. Appl.,
vol. 31, no. 2, pp. 240247, Mar./Apr. 1995.
Reiko Raute (S04M09) received the Dipl. Ing.
(FH) degree in electrical engineering degree from the
University of Applied Science Jena, Jena, Germany,
in 2005. From 2005 to 2009 he worked on his
Ph.D. degree project about sensorless control of ac
machines at the University of Malta, Msida, Malta.
He received the Ph.D. degree from the University of
Nottingham, Nottingham, U.K., in December 2009.
During his undergraduate studies, he visited the
University of Nottingham and the University of
Adelaide, Adelaide, Australia, working on electric
drive projects. He is currently with Carlo Gavazzi (Malta) Ltd.
Cedric Caruana (S01M05) received the Ph.D.
degree in electrical engineering from the University
of Nottingham, Nottingham, U.K., in 2003.
He was with Carlo Gavazzi (Malta) Ltd. and ST
Microelectronics (Malta) Ltd. as a Process Engineer
and Senior Test Engineer, respectively. He has been
with the Faculty of Engineering, University of Malta,
Msida, Malta, following receipt of the Ph.D. degree,
where he is currently a Senior Lecturer. His main
research interests are the control of ac drives, renew-
able energy conversion, electrical power systems,
and energy efciency.
Dr. Caruana is a member of the Institution of Engineering and
Technology, U.K.
1998 IEEE TRANSACTIONS ON INDUSTRY APPLICATIONS, VOL. 46, NO. 5, SEPTEMBER/OCTOBER 2010
Cyril Spiteri Staines (S94M98) received the
B.Eng.(Hons.) degree from the University of Malta,
Msida, Malta, in 1994, and the Ph.D. degree from
the University of Nottingham, Nottingham, U.K., in
1999, with his Ph.D. work on sensorless control of
ac drives carried out in the School of Electrical and
Electronic Engineering.
In 1995, he joined the University of Malta as an
Assistant Lecturer, where he became a Lecturer in
1999, Senior Lecturer in 2004, and Associate Profes-
sor in 2007, and he currently heads the Department
of Industrial Electrical Power Conversion. From 2003 to 2004, he was a
Postdoctoral Researcher and Visiting Lecturer at the University of Nottingham.
His research interests include sensorless ac motor drives and grid connection of
renewable energy sources, in particular, for wind-energy systems.
Prof. Staines is a member of the Institution of Engineering Technology, U.K.
Joseph Cilia received the Ph.D. degree in sensor-
less control of ac motors from the University of
Nottingham, Nottingham, U.K., in 1997.
He has been with the University of Malta, Msida,
Malta, where he was a Lecturer in power electronics
and drives, following receipt of the Ph.D. degree.
His research interests cover ac motor control and
encoderless techniques, power converters for motor
drive systems, battery management in motive-
power applications, grid-connected renewable en-
ergy sources, and energy efciency.
Mark Sumner (SM05) received the B.Eng. degree
in electrical and electronic engineering from Leeds
University, Leeds, U.K., in 1986, and the Ph.D.
degree in induction-motor drives from the University
of Nottingham, Nottingham, U.K., in 1990.
He was with Rolls Royce Ltd., Coventry, U.K.,
before embarking on his Ph.D. studies. After receipt
of the Ph.D. degree, he joined the University of
Nottingham as a Research Assistant, and where he
was appointed a Lecturer in October 1992, and is cur-
rently an Associate Professor and Reader in power
electronic systems. His research interests cover control of power electronic
systems, including sensorless motor drives, diagnostics and prognostics for
drive systems, power electronics for enhanced power quality, and novel power
system fault-location strategies.
Greg M. Asher (SM02F07) received the B.S. and
Ph.D. degrees in bond-graph structures and general
dynamic systems from Bath University, Bath, U.K.,
in 1976 and 1979, respectively.
He has been with the University of Nottingham,
Nottingham, U.K., where he was appointed a Lec-
turer in control in the School of Electrical and Elec-
tronic Engineering in 1984. He developed an interest
in motor drive systems, and was appointed Professor
of electrical drives in 2000, and is currently the
Associate Dean of the Engineering Faculty. He has
published over 200 research papers and has received over 5M in research
contracts.
Dr. Asher is an Associate Editor of the IEEE Industrial Electronics Society
(IES). He has been a member of the Executive Committee of the European
Power Electronics Association and Chair of the Power Electronics Technical
Committee of the IES.

Das könnte Ihnen auch gefallen