Sie sind auf Seite 1von 13

Mechanical Systems and Signal Processing 83 (2017) 228240

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Active magnetic bearings dynamic parameters identification


from experimental rotor unbalance response
Yuanping Xu a, Jin Zhou a,n, Long Di b, Chen Zhao a
a
College of Mechanical and Electrical Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing 210016, China
b
Rotating Machinery and Control Laboratory (ROMAC), University of Virginia, Charlottesville, VA 22904-4743, USA

a r t i c l e in f o abstract

Article history: Active magnetic bearings (AMBs) support rotors using electromagnetic force rather than
Received 14 July 2015 mechanical forces. It is necessary to accurately identify the AMBs force coefficients since
Received in revised form they play a critical role in the rotordynamic analysis including system stability, bending
9 May 2016
critical speeds and modes of vibrations.
Accepted 10 June 2016
Available online 29 June 2016
This paper proposes a rotor unbalance response based approach to identifying the
AMBs stiffness and damping coefficients during rotation. First, a Timoshenko beam finite
Keywords: element (FE) rotor model is created. Second, an identification procedure based on the FE
Active magnetic bearings model is proposed. Then based on the experimental rotor unbalance response data from
Dynamic parameters identification
1200 rpm to 30,000 rpm, the AMBs dynamic force parameters (stiffness and damping) are
Finite element method
obtained. Finally, the identified results are verified by comparing the estimated and ex-
Stiffness and damping
perimental rotor unbalance responses, which shows high accuracy.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Active magnetic bearings generate forces through magnetic fields rather than mechanical forces as in lubricated fluid
films or contact of rolling element bearings; therefore, the special advantage of AMBs is that there is no contact between
bearing and rotor, and this permits operation with no lubrication, no mechanical wear, long life, lower costs and high
attainable rotating speed [1,2]. Another attractive advantage of AMBs is that the dynamic force parameters, stiffness and
damping, are closely related to the feedback controller parameters, which can be changed easily [3], such that the ro-
tordynamics can be controlled and changed actively through the bearings.
For either AMBs or traditional mechanical bearings, it is vital to accurately obtain the dynamic force bearing parameters
since these coefficients for a rotor system are the foundation for the rotor dynamics analysis including system stability,
bending critical speeds, modes of vibrations, and response of rotating dynamic systems. The force coefficients of the tra-
ditional mechanical bearing are commonly modeled as stiffness and damping, which are also used for the AMBs. Compared
with traditional bearing system, the AMBs system is open loop unstable and feedback control is needed for levitation. Apart
from the rotor dynamics analysis, from the control design point of view, it is also vital to accurately predict the AMBs
dynamics parameters to optimize the control strategies design since the AMBs stiffness and damping are closely related to
the feedback controller parameters.
Humphris et al. [4] investigated the stiffness and damping properties of a magnetic bearing with two control algorithms

n
Corresponding author.
E-mail addresses: ypxu@nuaa.edu.cn (Y. Xu), zhj@nuaa.edu.cn (J. Zhou).

http://dx.doi.org/10.1016/j.ymssp.2016.06.009
0888-3270/& 2016 Elsevier Ltd. All rights reserved.
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 229

Front Sensor Rear Sensor

Rotor

Thrust Front Motor Rear Thrust


AMB AMB AMB AMB

Fig. 1. An overview of the rotor AMB test rig.

at static state; it was shown that varying amounts of position and velocity feedback affected the stiffness and damping
characteristics of the bearing, respectively. Williams et al. [5] proposed the estimated equivalent stiffness and damping
coefficients based on a theoretically derived frequency dependent feedback controller transfer function. However, the
theoretical estimated equivalent stiffness and damping coefficients procedures do not consider the time lags in the digital
control system. Lim et al. [3] experimentally obtained the stiffness and damping of a one-axis electromagnetic suspension
system under the shaker excitation. Similarly, under the shaker excitation, Lim [6] identified the parameter of five-degree-
of-freedom hybrid magnetic bearings for blood pump applications with proportional-integral-derivative (PID) controllers.
Tsai [7] applied the wavelet transform algorithm to identify the magnetic damping and magnetic stiffness coefficients.
Kozanecka [8] applied an unbalance excitation with the rotating vector force and the dynamic coefficients of the bearing
under nominal operating conditions are presented by calculating unbalance response, but did not consider the differential
location between sensors and actuator of AMBs. Although numerous studies have attempted to estimate these AMBs
coefficients directly or indirectly, few papers obtain AMBs coefficients under rotating condition. This paper seeks to address
this issue, by proposing a rotor unbalance response based approach to identifying the AMBs stiffness and damping coef-
ficients Fig. 1.
The identification method using unbalance response has been widely adopted in rolling element bearings and sliding
bearings [9,10], however few works have been reported on its application to AMBs. Compared with other identification
method such as under the shaker excitation, this method does not need additional devices and unbalance force excitation is
the simplest form of external excitation. More importantly, this method permits identification under the operating con-
dition. Before the parameters identification, a precise rotor model is indispensable. For the rotor adopted in the AMBs
system, its theoretical dynamic characteristics such as resonance and mode shape obtained from the FE model are not
always coincident with the experimental data since the rotor is often assembled by several components with interference
fits. For example, a typical rotor in AMBs system includes rotor shaft, front/rear AMB lamination stacks, thrust AMB

Fig. 2. Rotor of the AMB system.


230 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

lamination stacks, sensor rings and motor (Fig. 2), which are assembled with interference fits and shrink effects, con-
tributing to the bending stiffness of the rotor shaft. Previous work [11] by us has already detailed the rotor modeling, model
updating and updated model verification. Here, the updated FE rotor model is employed directly to calculate the AMBs
dynamic parameters. Before the test rig staring operation, we add two unbalance mass to the rotor. Then using the ex-
perimental unbalance response data, we identify the AMBs stiffness and damping coefficients during rotation ranging from
1200 rpm to 30,000 rpm. In order to make the results reliable, the verification by comparison experimental unbalance
response amplitude and phase data with the calculated unbalance response data using the identified coefficients is carried
out. The verification shows that this identified method is accurate and reliable.
The remainder of the paper is organized as follows. Section 2 describes the rotor AMBs test rig employed in this paper
and the mathematical modeling of AMBs rotor. Section 3 presents the identification method based on rotor unbalance
responses. Section 4 describes the unbalance responses and experimental results. Section 5 describes the verification of
identified results. Conclusions are drawn in Section 6.

2. Description of rotor AMBs test rig

2.1. Test rig description and instrumentation

The experimental test rig for this study is a five degree of freedoms (DOFs) rotor AMBs system designed and built as a
research platform at Nanjing University of Aeronautics and Astronautics, pictured in Fig. 1. The rotor is supported by two
radial and two thrust AMBs and is 0.468 m long and weighs around 2.4 kg. A 1.5 kW AC asynchronous induction motor rotor
is located in the middle of the rotor between the front AMB and rear AMB. Two laminated silicon steel sheets are mounted
at the end of rotor shaft for the two radical support AMBs and one laminated silicon steel sheet is assembled in the middle
for the driven motor. The specific details of the rotor are listed in Table 1. The radical and axial air gaps are 0.3 mm and
0.5 mm, respectively. Two rolling element ball bearings are assembled at the radical support AMB casings to prevent da-
mage to AMBs when the rotor is out of control.

2.2. Flexible rotor modeling

Most industrial rotors are flexible and an accurate rotor model is the fundamental of parameters identification. The rotor
model is often obtained by the finite element method. NelsonTimoshenko beam element is employed to simulate the rotor
system, and the gyroscopic moments, rotatory inertia, and shear deformation of the shaft are taken into consideration in
this study [12]. 53 NelsonTimoshenko beam finite element matrices are adopted to model the rotor according to the
geometrical and mass information, pictured in Fig. 3. The assembled parts such as lamination stacks are modeled as lumped
mass onto the corresponding nodes. Ignoring two DOFs in axial direction, each node contains 2 translational and 2 rota-
tional DOFs, therefore, the whole system possesses 216 DOFs. After assembling the governing equations for all the elements
and incorporating the boundary conditions, the linearized equations of motion for the shaft can be expressed as

Mq +( C+G) q+
Kq=f ( t ), (1)

where q and f (t ) are generalized displacement and generalized force vector in two radical directions; is the rotation
speed; M, C and K represent square symmetric mass, damping and stiffness matrices, respectively; G is the skew symmetric
gyroscopic matrix.
The FE code compiled in MATLAB for this rotor does not contain the unknown interactions of shrink fit interfaces,
inhomogeneous materials, small geometrical details, and so on, which cause potential errors to the modeling. In order to
obtain a relative accurate FE rotor model before identification, a precise FE rotor model is indispensable since any error in
the modeling of the flexible shaft will cause significant errors in the results. Previous work [11] by us has already detailed
updated the FE rotor model and validated it. For the rotor of our test rig in this study, its theoretical freefree mode shapes
plot without model updating is shown in Fig. 3. Table 2 lists the first forth bending mode modal frequencies before and after
model updating.

Table 1
Specific details of the rotor.

Label Name

A/L Front/rear thrust AMB lamination stacks


B/K Magnetic isolation ring
C/J Front/rear radical AMB lamination stacks
D/I Front/rear sensor ring
E Rotor shaft
F/H Motor rotor fix stacks
G Motor rotor
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 231

Fig. 3. Theoretical mode shapes of freefree rotor.

Table 2
Experimental and theoretical (no updating, updated) modal frequencies.

Bending Calculated freq. Measured freq. Updated freq.


modes (Hz) (Hz) (Hz)

1 454.9 477.5 477.5


2 1137.6 1167.5 1167.5
3 1802.7 1857.5 1855.4
4 2733.2 2805 2827.7

3. Method of AMBs dynamic parameters indentation

3.1. Theoretical equivalent stiffness and damping of one DOF AMB model

The AMB system is open loop unstable and feedback control is needed for levitation. Fig. 4 depicts a magnetic bearing
control loop system in one degree of freedom, which includes power amplifier, bearing electromagnet, controller, sensor
and rotor (flotor for non-rotating objects) [1].
The magnetic force for a single DOF AMB system can be expressed by its position stiffness Kyy and current stiffness Kiy
[1,4]:

power
amplifier

electromagnet

rotor / flotor

controller

sensor

Fig. 4. The basic magnetic bearing control loop.


232 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

Fig. 5. Block diagram of one DOF AMB system.

fm ( x,i)=Kyy x+Kiy i. (2)

The simplified control loop block diagram of one DOF AMB suspension system [5] is shown in Fig. 5, where G (s ) is the
feedback controller.
According to Fig. 5, the transfer function of one DOF AMB system can be written as

Y ( s) 1
= .
F ( s ) ms2+Kyy+Kiy G ( s ) (3)

Let s = j , where j = 1 , and the transfer function of Eq. (3) in the frequency domain can be expressed as

Y ( j) 1
= ,
F ( j) m2+Kyy+Kiy G ( j) (4)

where Y (j) is the Fourier transform of the displacement Y (s ); F (j) is the Fourier transform of excitation force F (s ). Ac-
cording to the Newton's law, the equation of motion of this magnetic bearing system can be expressed as

my +cy +ky=f y , (5)

where m, k and c represent the rotor mass, stiffness and damping coefficients, respectively. Using the Laplace transform
Eq. (5) can be written as

ms2Y ( s )+csY ( s )+kY ( s )=F ( s ). (6)

Let s = j , where j = 1 , and the Eq. (6) in the frequency domain can be expressed as

Y ( j) 1
= .
F ( j) m2+k+cj (7)

Based on Eqs. (4) and (7), we can acquire the equivalent stiffness and damping coefficients of the one DOF magnetic
bearing system, which is written as

k eq=Kyy+Re { Kiy G (s ) };
Im { Kiy G ( s ) }
ceq= .
(8)

If the control strategy is the proportional-integral-derivative (PID) algorithm, the controller transfer function G (s ) can be
written as

Ci
G ( s )=Cp+Cd s + ,
s (9)

the equivalent k and c coefficients can be expresses as follow:

k eq=Kiy Cp+Kyy;

C
ceq=Kiy Cd i2 .
(10)

keq and ceq are called equivalent coefficients because these coefficients are obtained from the analysis of the theoretical
frequency-dependent feedback controller transfer function G (s ), other than from the real experimental data. This analysis
method does not consider eddy current loss, hysteresis loss, ohmic loss and time delay in the digital signal processing,
amplifiers, feedback sensors etc. More importantly, for some robustness modern control techniques such as adaptive control
strategies, it is difficult to obtain a certain controller transfer function G (s ), which makes it difficult to get the real keq and ceq
coefficients, therefore, an experimental identification is indispensible.
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 233

3.2. Experimental stiffness and damping identification of rotor AMBs system

Since the theoretical equivalent stiffness and damping values are not accurate or difficult to acquire directly, estimating
these parameters through experimental data provides another reasonable solution. In the community of mechanical bearing
coefficient identification, various methods have been proposed. Here, based on the [13], we use unbalance responses to
calculate the AMBs stiffness and damping coefficients through online operation.
Here we use the updated FE rotor model to obtain experimental stiffness and damping coefficients. In Eq. (1),
Mq +( C+G) q +Kq = f ( t ), q and f are generalized displacement and generalized force vector in two radical directions, which
are written as

q= q1 qB1 qB2 qN ;
T

qi= xi yi xi yi ,

f = f1 fu fN ;
T

fi = fxi f yi Mxi Myi , i=1N . (11)

where qB1 and qB2 represent the displacements at the front, rear AMBs; xi and yi represent translations in the x and y
directions; xi and yi are the angular displacements about the y and x axes, respectively; fu is the unbalance excitation force
caused by unbalance mass, which can be written as



fx

cos ( t +)

real e (
i ( t + )
)

fy
fu =

=mu r2 sin ( t +) =mu r2 real( ie
i ( t + )
)
,
0 0
0
0 0
0 (12)

where mu is the unbalance mass; r is the radius between the unbalance location and the axis of shaft rotation. Note that the
periodic forced excitation with frequency caused by rotor unbalance is synchronous with rotating speed, i.e. = . In
general, f ( t )=Feit , so the unbalance response possesses the same frequency as the excitation, i.e. q = qeit . Therefore, Eq. (1)
can be written as the following form,
( KM2)+i ( C+G) q=F ,
(13)

and the transfer function between unbalance excitation and displacement (dynamic stiffness matrix) is

H = ( KM2)+i ( C+G) . (14)

The stiffness K and damping C matrices include the rotor coefficients (KR , CR ) and the AMBs coefficients (KB , CB ), therefore
K and C can be described as
K=KR+KB;

C=CR +CB, (15)

where
0 0

K b1 Cb1
KB= 0 ;CB= 0 .

K b2 Cb2
0 0 (16)

To simplify the calculation, ignoring cross-coupling effect, the AMBs stiffness and damping matrices incorporate the
following (yet unknown) coefficients:
Kxx 0 Cxx 0
K bi= ;Cbi= , i=1, 2.
0 Kyy i 0 Cyy i (17)

Therefore, let the Eq. (15) into Eq. (14), the transfer function Eq. (14) can be written as

H = ( KR+KBM2)+i ( CR +CB +G)


234 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

= ( KRM2)+i ( CR +G)+KB+iCB . (18)

Here, the dynamic rotor stiffness matrix HR and dynamic AMBs stiffness matrix HB can be defined as

HR= ( KRM2)+i ( CR +G) ,


HB=KB+iCB. (19)

Combining the dynamic rotor stiffness matrix and dynamic AMBs stiffness matrix together, the equation of motion of the
rotor AMB system can be written as below:

Hq=( HR+HB ) q=F. (20)

In order to identify the unknown AMBs coefficients, the algebraic system of Eq. (20) is reordered by use of matrix
operations to yield
0

HR q +HB q = 0 ,

F (21)

where the reordered displacement vector q is


ZB1

q = ZB2 ,

ZO (22)

where ZB1 and ZB2 are the transitional displacement vectors of AMBs supporter places, which is expressed by

ZB1= xB1 yB1 ;ZB2= xB2 yB2 ,


T T
(23)

where ( xB yB ) are the measured synchronous rotor responses (in complex form), ZO is the transitional displacement vectors
except AMBs supporter places.
In the FE rotor model, we have 54 nodes, therefore, the dimension of q , HR and HB is 216  216. The AMBs supporting
position nodes in the FE model are located at node 14 and 46. Using qi to denote each number in the original displacement
vector q . Therefore, the specific form of original q is

q= q1 q2 q3 q4 q216
T , (24)

where q1 x1; q2 y1; q3 x1;q4 y1.


Therefore, the specific reordered form of q can be written as

q = q53
q54
q181
q182
q1:52
q55:180
q183:216
T , (25)

Here q53, q54 , q181 and q182 are the xB1, yB1, xB2, yB2 respectively. So ZB1=[ q53 q54 ]T ; ZB2 [ q181 q182 ]T Z O [ q1:52 q55:180 q183:216 .
T

The recorded HR and HB use the same matrix operations.


In order to calculate conveniently, HR and HB are partitioned into sub matrices like below:
HR11 HR12 HR13 HB1 0 0

HR= HR21 HR22 HR23; HB= 0 HB2 0,
HR31 HR32 HR33 0 0 0 (26)

where HB1 and HB2 are the dynamic front, rear AMBs stiffness matrix, respectively, which are written as
Kxxi+iCxxi 0
HBi= ,i=1,2,
0 Kyyi+iCyyi (27)

such that Eq. (21) becomes:


HR11 HR12 HR13 ZB1 HB1 0 0 ZB1 0

HR21 HR22 HR23 ZB2 + 0 HB2 0 ZB2 = 0 .
HR31 HR32 HR33
ZO 0

0 0 ZO F (28)

In generally, for a rotor AMBs system, the sensors and actuators are not located at the same axial location, in Fig. 6.
Therefore, the measured rotor response is not the actual displacement at the AMBs. We need to transform the measured
responses to AMBs positions responses using the following relationship [14]:
ZB1 = a Z m1+AZ o;
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 235

B1 B2

S1 S2

AMB sensor sensor AMB


input output output input

Fig. 6. Different locations of AMBs actuator and sensor.

ZB2=bZ m2+BZ o, (29)

where Zm1 and Zm2 are the measured response vectors at the sensor locations near the AMBs, a and b are the known scalar
functions of the interpolation functions, and A and B are the matrix functions of the interpolation functions.
In Eq. (28), the motion of AMBs support places (ZB1,ZB2 ) and load vector F are obtained from measurements and un-
balance mass added to the rotor. So from the third row of this equation, ZO can be expressed as

{
ZO=HR331 F HR31ZB1HR32 ZB2 , } (30)

from the first and second row, we can obtain the follow equations:
HR11ZB1+HR12 ZB2+HR13 ZO= HB1ZB1;

HR21ZB1+HR22 ZB2+HR23 ZO= HB2 ZB2. (31)

Here, we define the AMBs transmitted forces fB1 and fB2, which can be written
fB1= (HR11ZB1+HR12 ZB2+HR13 ZO );
fB2 = ( HR21ZB1+HR22 ZB2+HR23 ZO ). (32)

Because ZO can be obtained from Eq. (30), the fB1 and fB2 values can be calculated. So the AMBs support coefficients are
determined by:
Kxx1+iCxx1 0
HB1= =f ZB11;
0 Kyy1+iCyy1 B1
Kxx2+iCxx2 0
HB2= =f ZB21.
0 Kyy2+iCyy2 B2 (33)

Note that, in the whole process of support coefficients calculation, we use the updated FE rotor model to provide an
accurate result.

4. Identification of AMBs stiffness and damping coefficients from unbalance responses

Before the unbalance response experiment, the rotor has been adjusted to nearly balanced condition under the dynamic
balancing machine to avoid the residual unbalance influence from the rotor itself.
The rotor unbalance response measurements are conducted for rotor speeds from 1200 rpm to 30,000 rpm, which is
above the first bending critical speed around 480 Hz (28,800 rpm). The unbalance response less than 1200 rpm is not
recorded since the unbalance displacement amplitude is small. Table 3 shows the unbalance mass distributions adopted in
this test. Two unbalance mass are added to rotor at node 17 and node 43, respectively. Fig. 7 depicts the unbalance mass

Table 3
Unbalance distributions on test rotor.

Added place Mass (g) Radius (mm) Phase (deg.)

Node 17 0.98 15 0
Node 43 0.78 15 180
236 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

Node 17

Mass added hole

Fig. 7. The unbalance mass added hole.

added hole at rotor node 17, and the unbalance mass added hole at rotor node 43 has the same structure.
During the whole unbalance test, the rotor displacements and rotating speed information are saved every 5 Hz
(300 rpm). Since the experimental data from the unbalance responses contains lots of noise, we need band pass filter to
process these signals. After storing the test data, the measured displacement values go through unbalance signal filter using
zero phase digital filters to help us extract the steady state amplitude and phase information. Then use one order Fourier
series based the least square method to fit the filtered signals, which could be written in the follow expression
y=a 0+a1 cos ( x)+b1 sin ( x), (34)

from the Eq. (34), the amplitude A and phase values under each rotating speed can be acquired though the parameters
a0, a1, b1, , which is written as Eq. (35)
2 2
A= a1 + b1
a1 .
=arctan
b1 (35)

Fig. 8 details the experimental unbalance responses data before and after filter under 100 Hz (6000 rpm) at sensor 1 in x
direction. Fig. 9 shows the unbalanced response test data process procedure.
Fig. 10 shows the measured unbalance responses at the sensor locations resulting from the unbalance masses under
rotating condition. Note that the measured rotor displacements are at sensor location radial planes and we need use Eq. (29)
to transform them from the measurements to AMBs actuators locations.
Fig. 11 shows the theoretical equivalent stiffness and damping coefficients calculated using the system transfer function
from Eq. (10) and the identified actual AMBs supporting stiffness and damping coefficients, obtained from the measured
rotor unbalance responses, using the method presented in this paper in Section 3. The theoretical equivalent coefficients
calculated from Eq. (10) of two AMBs are drawn using green line. Note that the control parameters are the same for both x
and y directions of each AMB, therefore only one green line indicates the theoretical values.
It is seen that before the first bending critical speed (around 480 Hz), the equivalent stiffness and damping values in-
creases steady with the running speed. Since the control parameters for both x and y direction of each AMB are the same,
the identified coefficients possess the same trend for the two orthogonal directions. However, there are still some small
unclose agreement, which may be attributed to the small discrepancy in mechanical and electrical performance. The

x 10
2

1
Amplitude (m)

-1

-2
1 1.01 1.02 1.03 1.04 1.05 1.06 1.07 1.08
Time (s)

Before filtering
After filtering

Fig. 8. Unbalance responses data under 100 Hz ( 6000 rpm) at sensor 1 in x direction.
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 237

Experimental unbalance
displacement data

Kaiser bandpass filter


design

Experimental rotating data


Zero phase digital filter

First order Fourier series First order Fourier series


fitting fitting

Fourier series function Fourier series function

Unbalance response Unbalance response


amplitude data phase data
Fig. 9. Flowchart of test data process procedure.

-5 -5
x 10 x 10
2 2
xm1
1.5 ym1 1.5
Amplitude (m)

Amplitude (m)

1 1
xm2

0.5 0.5 ym2

0 0
100 200 300 400 500 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)
200 200

100 100
Phase (degree)

xm1
Phase (degree)

ym1
0 0

xm2
-100 -100
ym2

-200 -200
100 200 300 400 500 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)

Fig. 10. Measured rotor unbalance responses (amplitude and phase) at sensor locations.

identified results fluctuate greater around the critical speed since the vibration near the circuital speed becomes greater. The
equivalent and test actual damping coefficients have the same trend although the test values fluctuate, however the dis-
crepancy for equivalent and test stiffness coefficients is great. The reason of these differences may attribute to that the
equivalent coefficients are obtained from the analysis of the theoretical frequency-dependent feedback controller transfer
function is simplified, ignoring the non-linearity in electromagnets, the eddy current loss, hysteresis loss, ohmic loss and
time delay in the digital signal processing, amplifiers, feedback sensors etc. Therefore the simplified theoretical calculation
cannot precisely stand the real AMBs stuffiness and damping and the experiment identification is indispensible.

5. Results verification

In order to make sure the identified values are accuracy, the verification is indispensible. We use these identified
coefficients as supporting parameters of the FE rotor model to calculate the unbalance responses under various operating
condition from 1200 rpm to 30,000 rpm. And then compare estimated the magnitude and phase with the measured during
the course of the experiment. Figs. 12 and 13 shows the comparisons between simulated and test amplitude and phase data
238 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

7.5 7.5
Kxx1
Stiffness (N/m) (log10)

Stiffness (N/m) (log10)


7 Kyy1 7
Kxx2
Keq1
6.5 6.5
Kyy2

6 6 Keq2

5.5 5.5

5 5
100 200 300 400 500 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)
6000 6000
Cxx1
5000 5000 Cxx2
Cyy1
Damping (N.s/m)

Damping (N.s/m)
4000 4000 Cyy2
Ceq1
Ceq2
3000 3000

2000 2000

1000 1000

0 0
100 200 300 400 500 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)

Fig. 11. Comparisons between effective and actual identified stiffness and damping coefficients. (For interpretation of the references to color in this figure,
the reader is referred to the web version of this article).

-5 -5
x 10 x 10
2 2

test data in xm1 test data in ym1


1.5 1.5
simulated data in xm1 simulated data in ym1
Amplitude (m)

Amplitude (m)

1 1

0.5 0.5

0 0
0 100 200 300 400 500 0 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)

50 200
test data in xm1
0 100
Phase (degree)

Phase (degree)

simulated data in xm1


test data in ym1
-50 0
simulated data in ym1

-100 -100

-150 -200
0 100 200 300 400 500 0 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)

Fig. 12. Comparisons between simulated and test amplitude and phase for front AMB.

for both AMBs. It can be observed that the unbalance response in the x and y axes from the finite element model agree with
the experimental measurements very well even around the critical speed (28,800 rpm, 480 Hz). Note that, we obtain the
same conclusion using the commercial finite element software to calculate the rotor unbalance responses under various
operating condition from 1200 rpm to 30,000 rpm. Here we just give the results calculated from FE rotor model created in
MATLAB in part 2.2, which has been updated and verified by our previous work. The verification results demonstrate that
the identification method proposed in this paper is effective and accurate.
Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240 239

-5 -5
x 10 x 10
2 2

1.5
Amplitude (m)

Amplitude (m)
1.5

1
test data in xm2 test data in ym2
1
0.5 simulated data in xm2 simulated data in xm2

0 0.5
0 100 200 300 400 500 0 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)
200 150

test data in ym2


100 100
Phase (degree)

Phase (degree)
simulated data in xm2

0 test data in xm2 50

simulated data in xm2


-100 0

-200 -50
0 100 200 300 400 500 0 100 200 300 400 500
Rotate speed (Hz) Rotate speed (Hz)

Fig. 13. Comparisons between simulated and test amplitude and phase for rear AMB.

6. Conclusion

This paper presented a supporting parameters identification strategy of AMBs under rotating condition. In order to
improve the identification accuracy, a Timoshenko beam FE rotor model is created. The proposed unbalance response based
identification method considers the differential location between sensors and AMBs. The identified supporting parameters
are verified by comparing the simulation and experimental rotor unbalance responses. The good consistency between es-
timated and experimental unbalance responses shows the proposed method is accurate and effective.
Compared with the theoretical analysis of control laws to obtain the stiffness and damping coefficients, unbalance mass
responses method is simply but reliable. The identified results show that the stiffness and damping coefficients on both x
and y axes vary along with the rotating speed, indicating that the magnetic force is affected by the operating speed. The
differences from the comparison between theoretical equivalent coefficients and experiential actual coefficients show that
the simplified theoretical calculation cannot accurately describe the real AMBs stuffiness and damping since various factors
are ignored in the idea condition.
Although the unbalance excitation is an easy and convenient excitation form for running rotor, some drawbacks cannot
be neglected. For example unbalances beyond certain level may generate excessive unbalance force, which may break or
damage the rotor system, especially for AMBs rotor system. In the experiment adopted in this paper, above 30,000 rpm
(500 Hz) rotating speed is not performed due to the safety reason. Hence, acquiring the dynamic parameters by impulse
response measurement or from the non-simplified system model may be a good choice. Current efforts are directed toward
the identification method mentioned above.

Acknowledgment

This research has been supported by Research and Innovation Program of Jiangsu Postgraduates (KYLX15_0296), National
Natural Science Foundation of China (51275240 and 51075200).

Appendix A. Supplementary material

Supplementary data associated with this article can be found in the online version at http://dx.doi.org/10.1016/j.ymssp.
2016.06.009.
240 Y. Xu et al. / Mechanical Systems and Signal Processing 83 (2017) 228240

References

[1] G. Schweitzer, E.H. Maslen, Magnetic Bearings: Theory, Design, and Application to Rotating Machinery, Springer, Berlin, Germany, 2009.
[2] G. Li, Z. Lin, P.E. Allaire, J. Luo, Modeling of a high speed rotor test rig with active magnetic bearings, J. Vib. Acoust. 128 (3) (2006) 269281.
[3] T.M. Lim, S.B. Cheng, Parameter estimation and statistical analysis on frequency-dependent active control forces, Mech. Syst. Signal Process. 21 (5)
(2007) 21122124.
[4] R.R. Humphris, R.D. Kelm, D.W. Lewis, P.E. Allaire, Effect of control algorithms on magnetic journal bearing properties, J. Eng. Gas Turbines Power 108
(4) (1986) 624632.
[5] R.D. Williams, F.J. Keith, P.E. Allaire, Digital control of active magnetic bearings, IEEE Trans. Ind. Electron. 37 (1) (1990) 1927.
[6] T.M. Lim, D. Zhang, J. Yang, S. Cheng, S.H. Low, L.P. Chua, X. Wu, Design and parameter estimation of hybrid magnetic bearings for blood pump
applications, Mech. Syst. Signal Process. 23 (7) (2009) 23522382.
[7] N.C. Tsai, H.Y. Li, C.C. Lin, C.W. Chiang, P.L. Wang, Identification of rod dynamics under influence of active magnetic bearing, Mechatronics 21 (6) (2011)
10131024.
[8] D. Kozanecka, Z. Kozanecki, T. Lech, Experimental identification of dynamic parameters for active magnetic bearings, J. Theor. Appl. Mech. 46 (1)
(2008) 4150.
[9] O.C. De Santiago, L. San Andrs, Field methods for identification of bearng support parameterspart I: identification from transient rotor dynamic
response due to impacts, J. Eng. Gas Turbines Power 129 (1) (2007) 205212.
[10] O.C. De Santiago, L. San Andrs, Field methods for identification of bearing support parameterspart II: identification from rotor dynamic response
due to imbalances, J. Eng. Gas Turbines Power 129 (1) (2007) 213219.
[11] Y. Xu, J. Zhou, L. Di, C. Zhao, Q. Guo, Active magnetic bearing rotor model updating using resonance and MAC error, Shock Vib. 2015 (2015).
[12] H.D. Nelson, A finite rotating shaft element using Timoshenko beam theory, J. Mech. Des. 102 (4) (1980) 793803.
[13] L. San Andres, O.C. De Santiago, Identification of bearing force coefficients from measurements of imbalance response of a flexible rotor, in: Pro-
ceedings of ASME Paper GT2004-54160.
[14] O.C. De Santiago, L. San Andrs, Experimental identification of bearing dynamic force coefficients in a flexible rotorfurther developments, Tribol.
Trans. 50 (1) (2007) 114126.

Das könnte Ihnen auch gefallen