Sie sind auf Seite 1von 10

IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO.

1, JANUARY 2013 129


Proposal and Analysis of a Novel Single-Drive
Bearingless Motor
Junichi Asama, Member, IEEE, Yuki Hamasaki, Takaaki Oiwa, and Akira Chiba, Fellow, IEEE
AbstractIn this paper, a novel design of a one degree of
freedom actively positioned bearingless motor, hereafter referred
to as a single-drive bearingless motor, has been proposed. The
concept is based on an axial ux motor where eld weakening
and strengthening with d-axis current modulation produce the
suspension force for axial motion control. The single-drive bear-
ingless motor requires neither additional windings, nor inverters
for noncontact magnetic suspension. It has only one three-phase
winding set and one three-phase inverter. To verify this novel
concept, theoretical calculations, nite-element method analysis,
and test machine development are carried out. The test results
successfully demonstrate the magnetic suspension and noncontact
rotation of the single-drive bearingless motor.
Index TermsAxial gap, bearingless motor, magnetic bear-
ing, magnetic levitation, one-axis, one-degree of freedom (DOF),
single-axis, single drive.
I. INTRODUCTION
A
MAGNETIC bearing is a mechanical element that can
suspend a rotating shaft without any mechanical contact
[1], [2]. The advantages of magnetic bearings are high-speed
and lubricant-free operation, no wear particles, no material
wear, and less heat generation. Therefore, magnetic bearings
have been applied to special machines such as turbomolecular
pumps, momentum and reaction wheels, machine tools, and
articial hearts. However, an additional motor element has to
be installed in the drive system to rotate the shaft; this causes
the axial length of the rotor shaft to become rather long.
To overcome the aforementioned problem, the bearingless
motor concept has been proposed since the late 1980s [2][4].
The bearingless motor can generate both suspension force
and torque in a single unit. To generate the suspension force,
additional windings are generally wound in the stator. The
bearingless motor combines a motor and magnetic bearings
together; thus, it possesses the advantages of compactness, a
simple structure, and cost reduction in comparison to a conven-
tional magnetic bearing drive.
Manuscript received July 31, 2011; revised November 6, 2011; accepted
December 21, 2011. Date of publication January 11, 2012; date of current
version September 6, 2012. This work was supported by grants from the
Mazda Foundation, the CASIO Science Promotion Foundation, and the NSK
Foundation for the Advancement of Mechatronics.
J. Asama, Y. Hamasaki, and T. Oiwa are with the Department of Mechanical
Engineering, Faculty of Engineering, Shizuoka University, Hamamatsu, 432-
8561, Japan (e-mail: tjasama@ipc.shizuoka.ac.jp; f0030048@ipc.shizuoka.
ac.jp; tmtooiw@ipc.shizuoka.ac.jp).
A. Chiba is with the Department of Electrical and Electronic Engineering,
Graduate School of Science and Engineering, Tokyo Institute of Technology,
Tokyo 152-8552, Japan (e-mail: chiba@ee.titech.ac.jp).
Digital Object Identier 10.1109/TIE.2012.2183840
Fig. 1. Schematic of four- and two-DOF controlled bearingless motors.
(a) Four-DOF control. (b) Two-DOF control.
To realize noncontact magnetic suspension, the rotor must
be magnetically stabilized in ve degrees of freedom (DOF),
including three-translational movements (x, y, z) and two-
tilting movements (
x
,
y
). The rotational speed (

z
) is also
regulated in magnetic bearings and bearingless motors. In these
research elds, the number of DOF for active positioning
control in ve DOF (x, y, z,
x
,
y
) excluding the rotation (
z
)
is conventionally discussed.
Various types of bearingless motors with active position
regulation of one to ve DOFs have been developed [3][24].
For example, a general four-DOF actively positioned bearing-
less motor with a cylindrical-shaped rotor and two bearing-
less motor units that are constructed in tandem is shown in
Fig. 1(a). Radial (x, y) and tilting (
x
,
y
) motions are actively
controlled. The bearingless motor units generate torque, and
thus, the rotational speed (

z
) is also regulated. An axial motion
(z) of the rotor is passively stabilized with passive magnetic
couplings between the rotor and the stator. In addition to a
motor drive inverter, two additional three-phase inverters are
required for the magnetic suspension. In a ve-DOF actively
positioned bearingless drive system, a thrust magnetic bearing
is additionally installed into the four-DOF-controlled bearing-
less motor for active positioning control of the rotor. For active
thrust motion control, an additional single-phase inverter is
required.
The increase in the number of actively positioned DOFs
causes increases in size, power consumption, and cost; ad-
ditionally, it complicates the system control. A two-DOF-
controlled bearingless motor has been proposed in [5]. Fig. 1(b)
shows a schematic diagram of an inner-disk-rotor-type bear-
ingless motor. The radial motion (x, y) of the rotor is ac-
tively regulated. The remaining DOFs, except the rotational
motion, are passively stabilized. The rotational motion (
z
) is
also actively controlled by regulating torque. To successfully
realize noncontact operation, two displacement sensors, two
sets of three-phase windings, and two three-phase inverters are
required in the two-DOF actively positioned bearingless motor.
Alternatively, two displacement sensors, one set of four- or
0278-0046/$31.00 2012 IEEE
130 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 1, JANUARY 2013
TABLE I
CLASSIFICATION OF BEARINGLESS DRIVES BY THE NUMBER OF DOF FOR ACTIVE POSITIONING
ve-phase windings, and one four- or ve-phase inverter are
required [23], [24].
To simplify the structure and the control system, an axial-
gap motor with one-DOF active position regulation has been
proposed for the bearingless drive [15], [16]. An axial force
is regulated to control the rotor axial position. A disk-shaped
rotor is sandwiched between two axial-gap stators. A three-
phase winding is wound around each stator, and thus, two
inverters are necessary. A simple structure was proposed in an
earlier report with one stator, where one three-phase inverter
controls torque and axial force [17]. Additional repulsive-type
permanent magnet bearings were installed in these motors to
passively restrict the radial and tilting motions. As a result,
the axial motion (z) and the rotational speed (

z
) are actively
controlled. However, complete magnetic levitation of the rotor
has not yet been realized in these motors.
Some researchers have successfully proposed single-DOF
actively positioned magnetic bearings [25][29]. In these test
machines, the magnetic suspension and rotating parts are sepa-
rated. Both single-phase and three-phased inverters are required
for axial force and torque regulation, respectively. Thus, two
inverters are necessary. Table I shows the classication of the
bearingless motors by the number of actively positioned DOFs
for magnetic suspension. It also shows the number of required
inverters. For example, in the ve-DOF active suspension, two
three-phase inverters are required for four-axis active suspen-
sion, and an additional three-phase inverter is required in the
motor torque generation. Moreover, one single-phase inverter
is needed for the Z-axis active positioning. In two-DOF active
suspension, two-sets of three-phase inverters are required. For
four- or ve-phase bearingless motors, one three-phase inverter
module, and one single-phase inverter module are required.
In this paper, a novel design of the single-DOF actively posi-
tioned bearingless motor is proposed, hereafter, it is referred to
as a single-drive bearingless motor. The single-drive bearing-
less motor can realize complete noncontact magnetic suspen-
sion and rotation employing only one three-phase inverter, as
shown in Table I. It requires neither additional windings nor
inverters. The axial motion (z) is actively regulated by eld
weakening and strengthening. The radial and tilting motions
are passively stabilized by the magnetic coupling between the
rotor and the stator. The rotational speed (

z
) is also actively
controlled by regulating torque. In Section II, the structure
and principle of suspension force generation is introduced. In
Fig. 2. Structure of the proposed single-drive bearingless motor.
Section III, a theoretical analysis based on mathematical for-
mulas is performed. The axial force and torque are theoretically
derived. In Section IV, nite-element method (FEM) analysis
and experiments are described. The test results demonstrate
successful magnetic noncontact rotation up to 1400 r/min and
verify the feasibility of the novel concept of the single-drive
bearingless motor. Section II and parts of Section IV were
presented at an IES-sponsored conference [30]. This paper
presents signicant improvements, including the theoretical
analysis (Section III), the control method, and additional results
of the FEM analysis and experimental tests (Section IV).
II. SINGLE-DRIVE BEARINGLESS MOTOR
A. Denition
The number of DOFs of the magnetic bearings and bearing-
less drives in active magnetic suspension is discussed. The term
one DOF has also been expressed in the literature as single
DOF, one-axis, single-axis, etc. The single-drive bearingless
motor has only one-DOF active magnetic suspension, typically
in the axial direction (z). In addition, only one three-phase drive
inverter is employed to suspend and rotate the rotor. This type
of bearingless motor is proposed as a single-drive bearingless
motor.
B. Motor Conguration
Fig. 2 shows the conguration of the proposed single-drive
bearingless motor. The Z-axis indicates the rotational axis and
the direction of the magnetic suspension force. Two axial-gap
types of surface permanent magnet motors are combined. The
rotor has a cylindrical shape since the two disk-shaped rotor
ASAMA et al.: PROPOSAL AND ANALYSIS OF A NOVEL SINGLE-DRIVE BEARINGLESS MOTOR 131
Fig. 3. Schematic diagram of suspension force generation. (a) Positive suspension force. (b) Negative suspension force.
Fig. 4. Schematic diagram of passive stabilization. (a) Radial direction. (b) Tilting direction.
irons are connected by a shaft. A permanent magnet ring is
attached on the rotor iron surface at both ends. A sensor target
for axial displacement detection is located at the center of the
rotor iron. A general three-phase winding used for both rotation
and magnetic suspension is wound around the stator. No addi-
tional windings are required for active magnetic suspension.
C. Principle of Magnetic Suspension
Fig. 3 shows a schematic diagram of suspension force gen-
eration in a single-drive bearingless motor. The bias magnetic
ux, which is induced by the permanent magnets, produces
magnetic couplings between the rotor and the stator. If the rotor
is centered in the axial direction, the magnetic attraction forces
are the same at both ends of the rotor. The axial stability of the
rotor is enabled by active positioning control with suspension
force regulation and axial displacement detection.
As shown in Fig. 3(a), when the magnetic ux is strength-
ened in the left side air gap, the magnetic ux densities in the
air gaps at both stators are unbalanced. This ux unbalance
generates a positive suspension force. In a similar manner,
the ux weakening generates a negative suspension force, as
shown in Fig. 3(b). In Section III, the axial suspension force is
theoretically calculated. In Section IV, the FEM analysis results
are shown.
D. Principle of Passive Stabilization
In the other four axes, except the Z-axis, passive magnetic
suspension is adopted. Fig. 4 shows the passive stability of
the rotor in the radial (x, y) and tilting (
x
,
y
) directions.
When the rotor moves in the radial direction, as a result of the
magnetic couplings between the rotor and the stator, a restoring
force is generated, as shown in Fig. 4(a). Thus, the radial motion
of the rotor is inherently stable and passively stabilized only
when the axial position is actively controlled. In a similar way,
when the rotor is tilted, these magnetic couplings produce a
restoring force, as shown in Fig. 4(b). For passive stability in
the tilting direction, the axial length of the rotor should be more
than twice the rotor radius [25].
III. THEORETICAL CALCULATION
A. Magneto-Motive-Force Distribution
In this section, the axial suspension force and torque of
the proposed single-drive bearingless motor are theoretically
derived. First, we focus on the left part of Fig. 2, which is
the conguration of an axial-gap surface-mounted permanent
magnet motor. To simplify the calculations, a two-pole and
two-phase winding conguration is considered. Fig. 5 shows
the arrangements of the two-pole rotor and two-pole two-phase
winding of the single-drive bearingless motor. The motor has
two-phase windings, denoted as 2a and 2b. In the following
section, the magneto-motive-force (MMF) distribution, and
the permeance are assumed to be derived from the magnetic
ux distribution in the air gap. From the ux distribution,
components in the inductance matrix are derived to calculate
the magnetic energy. The axial suspension force and torque
can be derived from the partial derivative of the magnetic en-
ergy. In the theoretical calculations, the following assumptions
are made.
a) MMF spatial distribution is sinusoidal.
b) Stator surfaces are smooth and thus stator slot harmonics
can be neglected.
c) Magnetic saturation is neglected.
d) Permeability of the iron core is innite. The air and the
permanent magnet permeabilities are equal to the space
permeability.
e) Rotor displacement is small with respect to the
electro-magnetic-gap between the rotor and stator iron
surfaces.
f) Iron losses and magnetic leakage uxes are neglected.
The MMF space distributions A
2a
and A
2b
for the two-pole
two-phase winding with unity current are written as
A
2a
=N
2a
cos
s
(1)
A
2b
=N
2b
sin
s
(2)
132 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 1, JANUARY 2013
Fig. 5. Simplied representation of the two-pole rotor and the two-phase
winding.
where N
2a
and N
2b
are the amplitudes of the MMF dis-
tribution fundamental components. The symbol
s
is the
angular position starting from the X-axis, as shown in Fig. 5.
The permanent magnet can be replaced with a coil excited with
dc current. The MMF distribution for the permanent magnet is
similarly written as
A
p
= N
p
cos(
s

m
) (3)
where N
p
and
m
are the amplitude of the MMF distribution
fundamental component and the rotating angle of the rotor,
respectively.
B. Permeance
Fig. 6 shows a 3-D model and an equivalent electrical circuit
of the axial-gap surface-mounted permanent magnet motor. In
the left-hand gure, a disk rotor iron core, a two-pole permanent
magnet ring, and a concave-shaped iron stator are shown. The
axial distance between the surfaces of the stator teeth and the
permanent magnet is l
g
. The axial length of the permanent
magnet is l
p
. The total distance of the electro-magnetic-gap g is
dened as the sum of l
g
and l
p
.
In the right-hand gure, an equivalent electrical circuit for
this model is drawn. The resistors represent the reluctances of
the air gap and the permanent magnet. The reluctance values
depend on the rotor axial-gap distance. The iron reluctances are
assumed to be zero. The dc voltage sources represent the MMF
of the permanent magnet. The voltage sources are not included
at
s
=
m
+/2 and
s
=
m
+ 3/2, because A
p
is zero
at these stator angles, as given by (3). The ground symbols
indicate that the iron magnetic potential is assumed to be zero.
The voltage V
p
indicates the magnetic potential of the stator.
In Fig. 6, 12 branches are shown to illustrate the calculation
concept; however, in the actual mathematical calculations an
innite number of branches is assumed.
The reluctances of the air gap R
g
and the permanent magnet
R
p
can be calculated using l
g
and l
p
as
R
g
=
l
g

g
S
R
p
=
l
p

p
S
(4)
where
g
and
p
are the permeabilities of the air and the per-
manent magnet, respectively, and S is the effective area of the
stator teeth. As described in the aforementioned assumption (d),

g
and
p
are equal to
0
. Thus, the total reluctance R
0
is
obtained by
R
0
= R
g
+R
p
=
l
g

0
S
+
l
p

0
S
=
g

0
S
. (5)
The inverse of the magnetic reluctance gives the electro-
magnetic-gap permeance as
P
0
=
1
R
0
=

0
S
g
. (6)
This equation indicates that the permeance P
0
is assumed to be
uniform along the
s
direction.
C. Magnetic Flux Distribution
The current owing through the branches in Fig. 6 corre-
sponds to the air-gap ux. The ux distribution for the perma-
nent magnet
p
is calculated using the equivalent circuit shown
in Fig. 6 as

p
= P
0
_
1
2
A
p
+V
p
_
. (7)
Since an integral of the ux
p
around the rotor surface is zero,
according to Gausss law, V
p
can be calculated as
V
p
=
1
2
_
2
0
P
0
A
p
d
s
_
2
0
P
0
d
s
=
N
p
_
2
0
cos(
s

m
)d
s
2
= 0.
(8)
We can see that the magnetic potential V
p
is zero. Therefore,
the ux distributions
p
is

p
=
1
2
P
0
A
p
. (9)
In a similar manner, the magnetic ux
2a
induced by the 2a
winding can be obtained from the equivalent electrical circuit
[2] as

2a
= P
0
_
1
2
A
2a
+V
2a
_
(10)
where
2a
and the voltage V
2a
are the magnetic ux induced
by the 2a winding and the magnetic potential of the rotor,
respectively. The voltage V
2a
can be calculated as
V
2a
=
_
2
0
P
0
A
2a
d
s
_
2
0
P
0
d
s
=
N
_
2
0
cos
s
d
s
2
= 0. (11)
From this equation, we can see that the magnetic potential V
2a
is also zero. Therefore, the air-gap ux distributions
2a
and

2b
produced by the 2a and 2b windings when excited by a
unity current are

2a
=
1
2
P
0
A
2a
(12)

2b
=
1
2
P
0
A
2b
. (13)
ASAMA et al.: PROPOSAL AND ANALYSIS OF A NOVEL SINGLE-DRIVE BEARINGLESS MOTOR 133
Fig. 6. Three-dimensional model and equivalent electrical circuit.
D. Inductance Matrix
The inductance matrix is derived using the air-gap ux
distributions. Let us suppose that the ux linkages of 2a, 2b,
and the equivalent permanent magnet coil windings are
2a
,

2b
, and
p
, respectively. The instantaneous currents of these
windings are i
2a
, i
2b
, and i
p
, respectively. The ux linkage and
current relationships are expressed in a matrix form as
_
_

2a

2b
_
_
=
_
_
L
p
M
2ap
M
2bp
M
2ap
L
2a
M
2a2b
M
2bp
M
2a2b
L
2b
_
_
_
_
i
p
i
2a
i
2b
_
_
= [L][i] (14)
where [L] is dened as a 3 3 inductance matrix and [i]
is the current vector. L and M indicate the self- and mutual
inductances, respectively. The inductances dened in the above
matrix can be derived by integrating the product of the air-gap
ux and the MMF distributions as
L
2a
=
1
2
2
_
0

2a
A
2a
d
s
=

0
SN
2
2a
4g
(15)
L
2b
=
1
2
2
_
0

2b
A
2b
d
s
=

0
SN
2
2b
4g
(16)
L
p
=
1
2
2
_
0

p
A
p
d
s
=

0
SN
2
p
4g
(17)
M
2a2b
=
1
2
2
_
0

2a
A
2b
d
s
= 0 (18)
M
2ap
=
1
2
2
_
0

2a
A
p
d
s
=

0
SN
2a
N
p
4g
cos
m
(19)
M
2bp
=
1
2
2
_
0

2b
A
p
d
s
=

0
SN
2b
N
p
4g
sin
m
. (20)
N
2a
and N
2b
are assumed to be equal, thus L
2a
and L
2b
can be
written as
L
2a
= L
2b
= L
2
=

0
SN
2
2a
4
1
g
= L

2
1
g
(21)
where L
2
, the self-inductance, is a function of the air-gap
distance g. We separate the variable g from L
2
using the
constant L

2
= (
0
SN
2
2a
)/4 with units Hm. In a similar way,
the inductances can be written as
L
p
=

0
SN
2
p
4
1
g
= L

p
1
g
(22)
M
2ap
=

0
SN
2a
N
p
4
1
g
cos
m
= M

1
g
cos
m
(23)
M
2bp
=

0
SN
2b
N
p
4
1
g
sin
m
= M

1
g
sin
m
(24)
where L

p
and M

are constants independent of the air-gap


distance.
E. Axial Suspension Force and Torque
The axial suspension force and torque can be derived using
the stored magnetic energy W
m
, which is given by
W
m
=
1
2
[i]
T
[L][i]. (25)
Expanding the matrix in (25) results in
W
m
=
1
2
L
p
i
2
p
+
1
2
L
2a
i
2
2a
+
1
2
L
2b
i
2
2b
+ M
2ap
i
2a
i
p
+M
2bp
i
2b
i
p
. (26)
Substituting (21)(24) into (26) yields
W
m
=
1
2g
__
i
2
2a
+i
2
2b
_
L

2
+i
2
p
L

p
+ 2i
p
M

(i
2a
cos
m
+i
2b
sin
m
)
_
. (27)
If the rotor is slightly displaced from the nominal position
g
0
toward the axial direction, then the following equation is
obtained:
g = g
0
z (28)
where z is the axial rotor displacement. The axial suspension
force F
m
can be obtained by substituting (28) into (27) and
taking the partial derivative of W
m
with respect to z as
F
m
=
W
m
z
=
1
2(g
0
z)
2
_
_
i
2
2a
+i
2
2b
_
L

2
+i
2
p
L

p
+ 2i
p
M

(i
2a
cos
m
+i
2b
sin
m
)
_
. (29)
134 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 1, JANUARY 2013
Furthermore, i
2a
and i
2b
are transformed into direct-axis
(d-axis) and quadrature-axis (q-axis) currents, i
d
and i
q
, by the
following matrix:
_
i
2a
i
2b
_
=
_
cos
m
sin
m
sin
m
cos
m
_ _
i
d
i
q
_
. (30)
Substituting (30) into (29) yields
F
m
=
__
i
2
d
+i
2
q
_
L

2
+i
2
p
L

p
+ 2i
d
i
p
M

_
2(g
0
z)
2
. (31)
In the right-hand part of the shaft shown in Fig. 2, an
iron core without winding is facing the permanent magnet.
In a similar manner, the attractive force F
0
can be derived
by substituting g = g
0
+z and i
d
= i
q
= 0 into the result-
ing F
m
as
F
0
=
i
2
p
L

p
2(g
0
+z)
2
. (32)
Therefore, the total axial force F
z
is obtained as
F
z
=F
m
F
0
=
_
i
2
d
L

2
+i
2
q
L

2
+i
2
p
L

p
+ 2i
d
i
p
M

_
2(g
0
z)
2

i
2
p
L

p
2(g
0
+z)
2
. (33)
Since z is much smaller than g
0
, the following equation can be
obtained:
1
(g
0
z)
2
=
1
g
2
0
(1 z/g
0
)
2

1
g
2
0
. (34)
In addition, if the rst term in (33) is much smaller than the
fourth term, the axial suspension force can be linearized as
follows:
F
z
=
1
2g
2
0
_
i
2
q
L

2
+ 2i
d
i
p
M

_
. (35)
In the second term in the parentheses, the axial suspension force
F
z
increases linearly when the d-axis current i
d
increases. From
the rst term, it is seen that the q-axis current i
q
may inuence
the axial suspension force, as discussed in the next section.
Taking the partial derivative of W
m
with respect to a small
rotating angle
m
yields torque T as
T =
W
m

m
=
i
p
M

g
{i
2b
cos
m
i
2a
sin
m
} . (36)
Furthermore, the d q current transformation provides the
following simplication:
T =
i
p
i
q
M

g
. (37)
The torque increases linearly as the q-axis current i
q
increases
and is not affected by the d-axis current i
d
.
Fig. 7. Calculated suspension force and toque. (a) Suspension force with
respect to suspension current i
d
. (b) Torque with respect to torque current i
q
.
IV. FEM ANALYSIS AND EXPERIMENT
A. FEM Analysis
To verify the proposed concept, 3-D FEM analysis is car-
ried out using commercially available software (JMAG, JSOL
Corporation, Japan). The axial length and the diameter of the
rotor are 75 mm and 25 mm, respectively. The axial thickness
and the radial width of the permanent magnet ring are 0.5 mm
and 2 mm, respectively. The air-gap distance from the perma-
nent magnet surface to the stator iron core is designed to be
0.6 mm. Therefore, the total electro-magnetic-gap is 1.1 mm.
The number of turns around one tooth is 147. The rated current
density is set to be 8 A/mm
2
, and thus the rated line current of
this machine is 0.16 A.
Fig. 7(a) and (b) shows the calculated axial suspension
force and torque, respectively. The horizontal axes indicate
the suspension current i
d
and torque current i
q
. When i
d
equals i
q
with the rated line current of 0.16 A, i
d
and i
q
are
both approximately 0.2 A. The suspension force and torque
calculated by the FEM are pulsated with respect to the rotor
angular position, although these are constant in mathematical
calculations, as shown in (35) and (37). The average is plotted
in Fig. 7. The force and torque variations are caused by the
difference of assumptions in theoretical calculations.
As shown in Fig. 7(a), the calculated suspension force in-
creases linearly with an increase in the suspension current of i
d
.
The gradient of the least squares approximation, as indicated
with a dotted line, is dened as the suspension force constant
K
i
and identied as K
i
= 3.5 N/A, when i
q
= 0 A and the
rated current is 0.2 A. The gradient of the linear plot derived
from the theoretical calculation (35) is K
i
= 3.1 N/A when
i
p
M

= 3.75 10
6
AHm and L

2
= 1.43 10
6
Hm. The
ASAMA et al.: PROPOSAL AND ANALYSIS OF A NOVEL SINGLE-DRIVE BEARINGLESS MOTOR 135
Fig. 8. Calculated stiffness in the passively stabilized radial and tilting direc-
tions with or without eld weakening (i
d
= 0.2 A). (a) Calculated restoring
force with respect to the radial displacement. (b) Calculated restoring torque
with respect to the tilting angle.
error between the theoretical calculation and the FEM analysis
is approximately 10%. When i
q
= 0.6 A, a small offset can
be observed in the suspension force, which is theoretically
indicated in the rst term in (35). Although i
q
inuences the
axial suspension force, limited inuence of the torque current i
q
on the suspension force is observed. Based on the FEM analysis
results, i
q
L

2
is signicantly small with respect to i
p
M

, i.e., the
permanent magnet MMF is signicantly high with respect to
the MMF originating from the winding current. Thus, the rst
term in (35) can be neglected in this single-drive bearingless
motor.
As shown in Fig. 7(b), the torque increases linearly as the
q-axis current increases. The gradient of the least squares
approximation, as indicated with a dotted line, is 3.0 mNm/A
when i
d
= 0 A and the rated current is 0.2 A. The gradient
of the linear plot derived from the theoretical calculation (37)
is 3.4 mNm/A. The error between the theoretical calculation
and the FEM analysis is approximately 10%. The suspension
current does not affect the torque performance, as theoretically
indicated in (37).
Fig. 8 shows the calculated restoring force and torque, which
act on the rotor with respect to the radial and tilting motions,
respectively. The inuence of eld weakening with the rated
suspension current of i
d
= 0.2 A is considered in FEM
calculation. As shown in Fig. 8(a) and (b), the gradients of the
linear approximation are the radial and tilting stiffness (K
r
and
K

) and are equal to K


r
= 2.4 N/mm and K

= 2.2 Nm/rad,
respectively. Although the radial and tilting stiffness decrease
by 10% with eld weakening, the calculated results show that
the radial and tilting motions are passively stabilized. The
calculated axial stiffness K
z
is negative. The axial motion
of the rotor is inherently unstable. Thus, active positioning
control is required in the axial direction to realize the magnetic
suspension of the rotor.
B. Control Method
Active positioning control is required in the axial direction.
The rotor motion equation of in the axial direction is given as
m z = F
z
. (38)
The suspension force F
z
is given by the MacLaurin expansion
in (33) with respect to i
d
and z as
F
z
= f
d
+K
i
i
d
+K
z
z (39)
where
f
d
=
i
2
q
L

2
2g
2
0
K
i
=
i
p
M

g
2
0
K
z
=
i
2
q
L

2
+ 2i
2
p
L

p
g
3
0
.
The variables f
d
, K
i
, and K
z
are disturbance, suspension
force constant, and stiffness in the axial direction, respectively.
Substituting (39) into (38) yields
m z K
z
z = K
i
i
d
+f
d
. (40)
We often encounter this form of the equation of motion [1],
[2]. Since this system has a negative stiffness, it is inherently
unstable as indicated in (40). An active positioning control is
required to stabilize the axial rotor motion. In this single-drive
bearingless motor, a proportional-integral-derivative (PID) con-
troller is adopted.
Fig. 9 shows the control system diagram of the single-drive
bearingless motor. The aforementioned PID feedback loop sta-
bilizes the axial rotor motion by regulating the d-axis current i
d
.
Positive and negative d-axis currents cause the strengthening
and weakening, respectively, of a magnetic ux in the air
gap. The resulting magnetic unbalance produces positive and
negative suspension forces in the axial direction. Rotational
torque is controlled by regulating the q-axis current i
q
. The PI
feedback loop is adopted to control the rotational speed.
The rotor displacement z in the axial direction is detected
by a displacement sensor. The error between the reference
and measurement is amplied by the PID controller, and the
d-axis reference current i

d
is generated. The PI speed regulation
controller generates the q-axis reference current i

q
. To regulate
the d- and q-axis currents, the PI feedback loops are adopted.
The d- and q-axis voltage commands are transformed into the
a- and b-axis voltage commands v

a
and v

b
. These are then
transformed into the three-phase voltage references v

u
, v

v
, and
v

w
. The three-phase voltage source inverter is controlled by
the three-phase voltage commands and supplies the suspension-
winding currents i
u
, i
v
, and i
w
. Both rotational torque and
magnetic suspension force are generated for regulating the
rotational speed regulation and an active Z-axis positioning by
regulating the q- and d-axis currents, respectively. Thus, the
single-drive bearingless motor can be driven by only one three-
phase inverter.
136 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 1, JANUARY 2013
Fig. 9. Control system diagram of the single-drive bearingless motor.
Fig. 10. Laboratory prototype of the single-drive bearingless motor.
C. Fabrication of Prototype
Fig. 10 shows a fabricated laboratory prototype of the single-
drive bearingless motor. The rotor and stator cores are made
of bulk soft iron. Four segmented permanent magnets with an
arc angle of 90

are used to construct a circular shape. The


permanent magnets are made of NdFeB with a residual ux
density of approximately 1.2 T. The rotor cores are press-tted
in a duralumin shaft. The rotor is initially displaced by 0.1 mm
from the center position and touched down before the magnetic
suspension control is activated. Thus, the movable range of the
rotor is 0.1 mm. Although the rotor expands by 0.1 mm with
a temperature rise of 60 K, the test machine can still be driven.
In addition, the movable range can be extended by adjusting the
bolt installed at the stator center, although the start-up current
is increased.
A short-pitch winding is used with a copper wire diameter
of 0.16 mm. An axial rotor motion is measured with an eddy-
current-type displacement sensor (PU-03A, AEC Corporation,
Japan). The rotating angular position is detected by two hall
sensors. A voltage source three-phase pulse-width-modulation
inverter with a dc voltage of 36 V is controlled by a digital
micro-processor (SH7047, Renesas Technology Corporation,
Japan). The sampling rates for the suspension control and
current control are 300 s and 100 s, respectively.
D. Experiment
The magnetic suspension control was successfully imple-
mented. Fig. 11 shows the axial displacement waveform of the
rotor at startup. The rotor is initially displaced by 0.1 mm
in the axial direction and touches down. When the magnetic
suspension control is activated, the rotor moves to the center
position. Fig. 12 shows the axial displacement at a rotational
Fig. 11. Axial displacement of the rotor at startup.
Fig. 12. Axial displacement of the rotor at 0 r/min.
Fig. 13. Axial displacement of the rotor against impulse.
speed of 0 r/min. The maximum vibration amplitude is 2 m,
i.e., small compared to 0.1 mm. Fig. 13 shows an impulse
response of the rotor in the axial direction. When the rotor is
magnetically levitated at 0 r/min, an impulse force is applied to
the rotor in the axial direction. Although the rotor is displaced
by 40 m by the impulse force, the rotor moves immediately to
the initial position after 50 ms. This shows that stable magnetic
suspension can be achieved.
Table II shows measured model parameters compared with
the FEM calculation results. The axial suspension force con-
stant K
i
is statically evaluated. The prototype machine is
vertically placed so that external loads can be applied to the
rotor in the axial direction. In position feedback, the axial
suspension force is generated in the opposite direction to the
external load. The suspension force is equal to the load because
an integral controller is included in the active feedback loop
and is proportional to the suspension current i
d
. The suspension
ASAMA et al.: PROPOSAL AND ANALYSIS OF A NOVEL SINGLE-DRIVE BEARINGLESS MOTOR 137
TABLE II
COMPARISON OF THE MODEL PARAMETERS BETWEEN
MEASUREMENT AND FEM CALCULATION
Fig. 14. Axial displacement of the rotor at 1400 r/min.
force constant is measured when the rotating angular position
is changed. The average is K
i
= 3.7 N/A. An error between the
measurement and the FEM calculation is approximately 6%.
The radial stiffness K
r
is statically evaluated. The prototype
machine is horizontally placed so that external loads can be
applied in the radial direction of the rotor. The rotor radial
displacements against the various loads are measured when the
rotating angular position is changed. The average of the mea-
sured radial stiffness is K
r
= 2.2 N/mm. The tilting stiffness
K

is dynamically evaluated. When the rotor is magnetically


suspended, an impulse response in the tilting direction is mea-
sured. A tilting resonant frequency is identied as 29.8 Hz.
The tilting stiffness is calculated as K

= 1.8 Nm/rad using


the moment of inertial around the radial direction of 5.2
10
5
kgm
2
. Errors of the radial and tilting stiffness between
the measurement and the FEM calculation are approximately
10% and 20%, respectively.
The rotor is successfully driven without any mechanical
contact. Fig. 14 shows the axial displacement of the rotor
at a rotational speed of 1400 r/min. The maximum speed is
1400 r/min, and it is limited by the tilting vibration of the
rotor. The axial rotor vibration is increased when the rotational
speed increases. The peak-to-peak value of the rotor vibra-
tion is 37 m, which is smaller than the initial clearance of
0.1 mm. The experimental results demonstrate the realization
and the feasibility of the proposed novel concept of the single-
drive bearingless motor. At this stage of the research project,
no special effort is being devoted to damping improvement,
vibration suppression, and torque enhancement. These issues
will be addressed in the future. The single-drive bearingless
motor will be applied to centrifugal/axial ow pumps and
reaction or momentum wheels, as well as actuators used in a
clean environment.
V. CONCLUSION
A novel design for a single-drive bearingless motor is pro-
posed in this paper. The single-drive bearingless motor is de-
ned as a one-DOF actively positioned bearingless motor which
is driven by only one three-phase inverter. The axial suspension
force and torque are theoretically calculated. FEM calculations
show that the axial suspension force and torque are linearly
regulated by i
d
and i
q
, respectively. The test results demonstrate
successful magnetic non-contact rotation up to 1400 r/min and
verify the feasibility of the novel concept of the single-drive
bearingless motor.
REFERENCES
[1] H. BlEuler, M. Cole, P. Keogh, R. Larsonneur, E. Maslen, R. Nordmann,
Y. Okada, G. Schweitzer, and A. Traxler, Magnetic Bearings. Berlin,
Germany: Springer-Verlag, Nov. 2009.
[2] A. Chiba, T. Fukao, O. Ichikawa, M. Oshima, M. Takemoto, and
D. G. Dorrell, Magnetic Bearings and Bearingless Drives. Amsterdam,
The Netherlands: Elsevier, Mar. 2005.
[3] A. Chiba and T. Fukao, Electric rotating machinery with radial position
control windings and its rotor radial position controller, Japan Patent
2 835 522, application date 1989/1/18, issued date 1998/10/9.
[4] A. Chiba, T. Deido, T. Fukao, and M. A. Rahman, An analysis of bear-
ingless ac motors, IEEE Trans. Energy Convers., vol. 9, no. 1, pp. 6168,
Mar. 1994.
[5] R. Schb and N. Barletta, Principle and application of a bearingless slice
motor, JSME Int. J. Ser. C, vol. 40, no. 4, pp. 593598, 1997.
[6] T. Reichert, T. Nussbaumer, and J. W. Kolar, Bearingless 300 W PMSM
for bioreactor mixing, IEEE Trans. Ind. Electron., vol. 59, no. 3,
pp. 13761388, Mar. 2012.
[7] B. Warberger, R. Kaelin, T. Nussbaumer, and J. W. Kolar, 50 Nm/2500 W
bearingless motor for high-purity pharmaceutical mixing, IEEE Trans.
Ind. Electron., vol. 59, no. 5, pp. 22362247, 2012.
[8] E. F. Rodriguez and J. A. Santisteban, An improved control system for
a split winding bearingless induction motor, IEEE Trans. Ind. Electron.,
vol. 58, no. 8, pp. 34013408, Aug. 2011.
[9] X. Wang, Q. Zhong, Z. Deng, and S. Yue, Current-controlled multi-
phase slice permanent magnetic bearingless motors with open-circuited
phases: Fault-tolerant controllability and its verication, IEEE Trans.
Ind. Electron., vol. 59, no. 5, pp. 20592072, 2012.
[10] C. Li and W. Hofmann, Speed regulation technique of one bearingless
8/6 switched reluctance motor with simpler single winding structure,
IEEE Trans. Ind. Electron., vol. 59, no. 6, pp. 25922600, 2012.
[11] Z. Ren and L. S. Stephens, Force characteristics and gain determination
for a slotless self-bearing motor, IEEE Trans. Magn., vol. 42, no. 7,
pp. 18491860, Jul. 2006.
[12] W. K. S. Khoo, Bridge congured winding for polyphase self-bearing
machines, IEEE Trans. Magn., vol. 41, no. 4, pp. 12891295, Apr. 2005.
[13] S. M. Yang and C. L. Lin, Levitation and torque control of a PM syn-
chronous self-bearing motor with a single set of windings, in Proc. 33th
IEEE Ind. Electron. Soc., Nov. 2007, pp. 10331037.
[14] M. Ooshima, A. Chiba, T. Fukao, and M. A. Rahman, Design and
analysis of permanent magnet-type bearingless motors, IEEE Trans. Ind.
Electron., vol. 43, no. 2, pp. 292299, Apr. 1996.
[15] Q. D. Nguen and S. Ueno, Analysis and control of nonsalient permanent
magnet axial gap self-bearing motor, IEEE Trans. Ind. Electron., vol. 58,
no. 7, pp. 26442652, Jul. 2011.
[16] Q. D. Nguen and S. Ueno, Modeling and control of salient-pole perma-
nent magnet axial-gap self-bearing motor, IEEE/ASME Trans. Mecha-
tronics, vol. 16, no. 3, pp. 518526, Jun. 2011.
[17] S. Ueno and Y. Okada, Vector control of an induction type axial gap
combined motor-bearing, in Proc. IEEE/ASME Int. Conf. Adv. Intell.
Mechatronics, Atlanta, GA, Sep. 1923, 1999, pp. 794799.
[18] J. Asama, R. Nakamura, H. Sugimoto, and A. Chiba, Evaluation of
magnetic suspension performance in a multi-consequent-pole bearingless
motor, IEEE Trans. Magn., vol. 47, no. 10, pp. 42624265, Oct. 2011.
[19] J. Asama, R. Kawata, T. Tamura, T. Oiwa, and A. Chiba, Reduction of
force interference and. performance improvement of a consequent-pole
bearingless motor, Precis. Eng., vol. 36, no. 1, pp. 1018, Jan. 2012.
[20] J. Asama, M. Amada, N. Tanabe, N. Miyamoto, A. Chiba, S. Iwasaki,
M. Takemoto, T. Fukao, and M. A. Rahman, Evaluation of a bearingless
PM motor with wide magnetic gaps, IEEE Trans. Energy Convers.,
vol. 25, no. 4, pp. 957964, Dec. 2010.
[21] R. L. A. Ribeiro, F. E. F. Castro, A. O. Salazar, and A. L. Maitelli,
A suitable current control strategy for split-phase bearingless three-
phase induction machine, in Proc. IEEE Power Electron. Spec. Conf.,
Jun. 2005, pp. 701706.
[22] A. Chiba, K. Sotome, Y. Iiyama, and M. A. Rahman, A novel
middle-point-current-injection-type bearingless PM synchronous motor
138 IEEE TRANSACTIONS ON INDUSTRIAL ELECTRONICS, VOL. 60, NO. 1, JANUARY 2013
for vibration suppression, IEEE Trans. Ind. Appl., vol. 47, no. 4,
pp. 17001706, Jul./Aug. 2011.
[23] M. T. Bartholet, T. Nussbaumer, S. Silber, and J. W. Kolar, Compara-
tive evaluation of polyphase bearingless slice motors for uid-handling
applications, IEEE Trans. Ind. Appl., vol. 45, no. 5, pp. 18211830,
Sep./Oct. 2009.
[24] W. Gruber, W. Amrhein, and M. Haslmayr, Bearingless segment motor
with ve stator elementsDesign and optimization, IEEE Trans. Ind.
Appl., vol. 45, no. 4, pp. 13011308, Jul./Aug. 2009.
[25] I. D. Silva and O. Horikawa, An attraction-type magnetic bearing with
control in a single direction, IEEE Trans. Ind. Appl., vol. 36, no. 4,
pp. 11381142, Jul./Aug. 2000.
[26] J. Kuroki, T. Shinshi, L. Li, and A. Shimokohbe, Miniaturization of
a one-axis-controlled magnetic bearing, Precis. Eng., vol. 29, no. 2,
pp. 208218, Apr. 2005.
[27] A. Yumoto, T. Shinshi, X. Zhang, H. Tachikawa, and A. Shimokohbe,
A one-DOF controlled magnetic bearing for compact centrifugal blood
pumps, in Motion and Vibration Control. New York: Springer
Science+Business Media B.V., 2009, pp. 357366.
[28] T. Merkel, A. Arndt, J. Hoffmann, P. Nsser, K. Graichen, W. Neumann,
and J. Mller, Magnetic bearing in INCOR axial blood ow pump acts
as multifunctional sensor, in Proc. 9th Int. Symp. Magn. Bearings, Lex-
ington, KY, 2004.
[29] S. M. Yang and M. S. Huang, Design and implementation of a magneti-
cally levitated single-axis controlled axial blood pump, IEEE Trans. Ind.
Electron., vol. 56, no. 6, pp. 22132219, Jun. 2009.
[30] J. Asama, Y. Hamasaki, T. Oiwa, and A. Chiba, A novel concept of a
single-drive bearingless motor, in Proc. IEEE IEMDC, Niagara Falls,
ON, Canada, May 1518, 2011, pp. 642646.
Junichi Asama (M08) was born in Niigata, Japan,
in 1979. He received the B.S., M.S., and Ph.D.
degrees in mechanical engineering from the Tokyo
Institute of Technology, Tokyo, Japan, in 2002, 2004,
and 2006, respectively.
In 2006, he was a Postdoctoral Researcher in the
Precision and Intelligence Laboratory, Tokyo Insti-
tute of Technology. In 2007, he joined the Tokyo
University of Science as a Research Associate in the
Department of Electrical Engineering in the Faculty
of Science and Technology. In 2009, he joined the
Faculty of Engineering, Department of Mechanical Engineering, Shizuoka
University, Hamamatsu, Japan. He is engaged in research on bearingless motor
drive systems and their applications.
Dr. Asama is a member of the Institute of Electrical Engineers of Japan, the
Japan Society for Precision Engineering, and the Japan Society of Mechanical
Engineers in Japan.
Yuki Hamasaki was born in Shizuoka, Japan, in
1987. He received the B.S. degree from the De-
partment of Mechanical Engineering in the Faculty
of Engineering, Shizuoka University, Hamamatsu,
Japan, in 2010.
He is currently engaged in a single-drive bearing-
less motor project.
Takaaki Oiwa was born in Toyohashi, Japan, 1959.
He received the B.S., M.S., and Ph.D. degrees in the
Nagaoka University of Technology, Nagaoka, Japan,
in 1982, 1984, and 1993, respectively.
He is currently a Professor in the Faculty of Engi-
neering at the Shizuoka University. Since 1980, his
major research interests have been precision machine
system including precision mechanisms, precision
measurement.
Dr. Oiwa is a member of the Japan Society of Me-
chanical Engineers, the Japan Society for Precision
Engineering, and the International Federation for the Promotion of Mechanism
and Machine Science.
Akira Chiba (S82M88SM97F07) was born
in Tokyo, Japan, in 1960. He received the B.S., MS.,
and Ph.D. degrees in electrical engineering from
the Tokyo Institute of Technology, Tokyo, Japan, in
1983, 1985, and 1988, respectively.
In 1988, he joined the Tokyo University of Science
as a Research Associate in the Department of Elec-
trical Engineering in the Faculty of Science and
Technology. From 1992 to 1993, from 1993 to 1997,
from 1997 to 2004, and from 2004 to 2010 he
was Research Lecturer, Senior Lecturer, Associate
Professor and Professor, respectively. Since 2010, he has been Professor in the
Department of Electrical & Electronic Engineering in the Graduate School of
Science and Engineering in the Tokyo Institute of Technology. In 19901991,
he was a Natural Science and Engineering Research Council of Canada Inter-
national Postdoctoral Fellow in the Memorial University of Newfoundland,
Canada. He has been studying magnetically suspended bearingless ac motors,
super high-speed motor drives, and rare-earth-free-motors for hybrid and pure
electrical vehicles. He has so far published more than 817 papers including the
rst book on Magnetic bearings and bearingless drives in 2005 and submitted
58 patents.
Dr. Chiba received the IEEJ Paper Awards in 1998 and 2005. He received
the First Prize Paper Award from the Electrical Machine Committee in 2011.
He has been served as Secretary, vice-Chair, vice-Chair-Chair-elect, in the
Motor Sub-Committee in the Electric Machinery Committee in IEEE PES
in 20072008, 20092010, and 2011, respectively. He served as vice-Chair,
and Chair in the IEEE IAS Japan chapter in 20082009 and 20102011,
respectively. He has been a member of the Electric Machine Committee and the
Industrial Drives Committee in IEEE IAS. He is a member of IEEJ in Japan.

Das könnte Ihnen auch gefallen